2,023
Views
2
CrossRef citations to date
0
Altmetric
Review Article

Regulatory mechanisms of exopolysaccharide synthesis and biofilm formation in Streptococcus mutans

, , , , , , , & ORCID Icon show all
Article: 2225257 | Received 21 Nov 2022, Accepted 08 Jun 2023, Published online: 18 Jun 2023

ABSTRACT

Background

Dental caries is a chronic, multifactorial and biofilm-mediated oral bacterial infection affecting almost every age group and every geographical region. Streptococcus mutans is considered an important pathogen responsible for the initiation and development of dental caries. It produces exopolysaccharides in situ to promote the colonization of cariogenic bacteria and coordinate dental biofilm development.

Objective

The understanding of the regulatory mechanism of S. mutans biofilm formation can provide a theoretical basis for the prevention and treatment of caries.

Design

At present, an increasing number of studies have identified many regulatory systems in S. mutans that regulate biofilm formation, including second messengers (e.g. c-di-AMP, Ap4A), transcription factors (e.g. EpsR, RcrR, StsR, AhrC, FruR), two-component systems (e.g. CovR, VicR), small RNA (including sRNA0426, srn92532, and srn133489), acetylation modifications (e.g. ActG), CRISPR-associated proteins (e.g. Cas3), PTS systems (e.g. EIIAB), quorum-sensing signaling system (e.g. LuxS), enzymes (including Dex, YidC, CopZ, EzrA, lmrB, SprV, RecA, PdxR, MurI) and small-molecule metabolites.

Results

This review summarizes the recent progress in the molecular regulatory mechanisms of exopolysaccharides synthesis and biofilm formation in S. mutans.

Introduction

Dental caries is a common chronic infectious disease caused by bacteria that seriously endangers the oral and systemic health [Citation1,Citation2]. According to the Global Burden of Disease (GBD) 2019 study, the caries rate of permanent teeth ranks first among 369 major diseases [Citation3]. An estimated 2.0 billion (95% uncertainty interval, 1.8 to 2.3 billion) people with untreated caries of permanent teeth are present worldwide, and an age-standardized prevalence of 25,625.5 (22,281.1 to 29,372.0) per 100,000 person/year in 2019 [Citation4,Citation5]. Additionally, caries in deciduous teeth ranks first among all disorders in children aged 0 to 14 years in 2019, with 0.5 billion (0.4 to 1.6 billion) cases of caries in deciduous teeth and an age-standardized prevalence of 7,672.9 (6,113.1 to 9,339.6) [Citation4,Citation5]. S. mutans is closely related to the initiation and development of caries due to its aciduric and acidic properties, and ability to synthetize glucan to form biofilm [Citation6]. The ability to synthesize large amounts of glucan to form biofilm contributes to the permanent colonization of tooth surfaces by S. mutans and the local development of the extracellular polymer matrix [Citation7].

S. mutans biofilm formation is regulated by two independent mechanisms: sucrose-dependent and sucrose-independent [Citation8,Citation9]. In the absence of sucrose or early stages of adhesion, the interaction of antigen I/II (also known as P1, SpaP, Ag I/II, or PAc) with agglutinins in the saliva mediates the adhesion of S. mutans to the dental surfaces and bacterial aggregation [Citation7,Citation10–12]. In the presence of sucrose, glucosyltransferases secreted by S. mutans, including GtfB, GtfC and GtfD, utilizes the glucose part of sucrose as a substrate to form the growing polymer of glucan, which is also called exopolysaccharides [Citation7,Citation13]. The presence of exopolysaccharides enhances the strength of the local adhesion of S. mutans to the tooth surfaces [Citation14]. Moreover, the continuous production of exopolysaccharides in situ further expanded the three-dimensional matrix while forming a bacterial cell core wrapped by exopolysaccharides [Citation15]. It also provides a scaffold for the colonization of other microorganisms (e.g. Candida albicans) and the adhesion of other substances (including eDNA, lipoteichoic acids, proteins), further leading to biofilm formation [Citation16].

This review summarizes the molecular regulatory mechanisms affecting the formation of S. mutans biofilm. The signals that regulate the formation of S. mutans biofilm are considered from five different aspects: (1) upstream signals, such as second messengers; (2) transcriptional level, such as transcription factors and two-component systems; (3) post-transcriptional level, such as small RNAs; (4) post-translational level, such as acetylation, malonylation; (5) others, such as CRISPR-related proteins, PTS systems, quorum-sensing signaling system, enzymes and small-molecule metabolites ().

Figure 1.  The regulatory mechanisms of exopolysaccharides synthesis and biofilm formation in S. mutans. c-di-AMP controls S. mutans biofilm formation by regulating gtfB expression through the binding of CabPA to VicR, while Ap4A represses the expression of gtfs through some unknown pathways. VicR, EpsR and CovR regulate S. mutans exopolysaccharides synthesis and biofilm formation directly by binding to the promoters of gtfB and gtfC. Post-translational modifications are also involved in the regulation of Gtfs activities and biofilm formation in S. mutans.

Figure 1.  The regulatory mechanisms of exopolysaccharides synthesis and biofilm formation in S. mutans. c-di-AMP controls S. mutans biofilm formation by regulating gtfB expression through the binding of CabPA to VicR, while Ap4A represses the expression of gtfs through some unknown pathways. VicR, EpsR and CovR regulate S. mutans exopolysaccharides synthesis and biofilm formation directly by binding to the promoters of gtfB and gtfC. Post-translational modifications are also involved in the regulation of Gtfs activities and biofilm formation in S. mutans.

Regulatory formation of S. mutans biofilm from upstream signals

Second messengers

Bacteria use second messengers to conduct signals in response to the rapid change of the environment. In recent years, many nucleotide molecules or second messengers have been found as regulators of various physiological activities of bacteria, including biofilm formation [Citation17].

c-di-AMP

c-di-AMP is an important and ubiquitous second messenger in bacterial signaling, whose intracellular levels are regulated by the enzymes diadenylate cyclase (DAC) and phosphodiesterase (PDE) [Citation18]. DAC contains a highly conserved DAC/DisA_N domain, which utilizes two molecules of ATP or ADP as a substrate to synthesize c-di-AMP [Citation19,Citation20]. PDE contains a DHH/DHHA1 (Asp-His-His)/HD (Asp-His) domain, which breaks down c-di-AMP into two AMP molecules or 5′-phosphoadenylyl-(3′−5′)-adenosine (5’-pApA) [Citation21,Citation22]. Therefore, the DAC and PDE processes keep the intracellular c-di-AMP level within an appropriate range.

The DAC and PDE enzymes in S. mutans are encoded by cdaA and pdeA, respectively. Cheng et al. found that the c-di-AMP levels of S. mutans decreased significantly after knockout cdaA [Citation23]. Genes with increased expression were clustered in cellular polysaccharide biosynthetic processes, including gtfB, gtfC [Citation23]. Besides, both GTF enzyme activity and their expression at the gene level were significantly increased. Nevertheless, Peng et al. constructed a similar in-frame deletion mutant obtaining different results [Citation24]. Indeed, DAC deficiency significantly reduces the expression of gtfB and inhibits the synthesis of exopolysaccharides, but does not markedly affect the expression of gtfC and gtfD. Therefore, the involvement of c-di-AMP in the formation of S. mutans biofilms occurs through the regulation of gtfB expression. In addition, Peng et al. revealed that CabpA and CabpB are two receptor proteins for c-di-AMP [Citation25]. CabPA interacts with VicR, a response regulator that regulates gtfB expression, thereby promoting biofilm formation. These results demonstrate that c-di-AMP is an upstream signal that regulates biofilm formation in S. mutans.

Ap4A

Ap4A is a dinucleotide metabolite that consists of two adenosines joined in the 5′−5′ linkage by four phosphates [Citation26,Citation27]. During translation, aminoacyl-tRNA synthetases (AARS) catalyze the transfer of the AMP moiety from the aminoacyl adenylate to the ATP to form Ap4A in the absence of tRNA [Citation28]. Ap4A is symmetrically decomposed into two molecules of ADP or asymmetrically into AMP and ATP by the Ap4A hydrolase YqeK, which was identified and characterized in S. mutans [Citation29]. Zheng et al. showed that the in-frame deletion of the yqeK gene in S. mutans causes an increase in the intracellular Ap4A levels, while biofilm formation and water-insoluble exopolysaccharides production are decreased [Citation30]. The knockout of the yqeK gene leads to the downregulation of important virulence genes related to biofilms, such as gtfB, gtfC, and gtfD [Citation30]. Although the mechanisms involved in the role of Ap4A need to be further investigated, these results suggest that it affects biofilm formation in S. mutans by affecting the expression of gtfs and the activity of Gtfs.

Regulation of S. mutans biofilm formation at the transcriptional level

Transcription factors

Some transcription factors also play a very important role in the formation of S. mutans biofilm. The exopolysaccharides synthesis regulator (epsR) gene is one of the transcription factors of the multiple antibiotic resistance regulator (MarR) family in S. mutans. Chen et al. demonstrate that EpsR specifically binds to the promoter regions of gtfB, consequently negatively regulating gtfB expression and exopolysaccharides production in S. mutans [Citation31].

RcrR, a rel competence-related regulator, is a transcription factor belonging to the multiple antibiotic resistance regulator (MarR) family in S. mutans. RcrR specifically binds to the promoter regions of Smu.1185 and Smu.2038 to regulate the carbohydrate transport of mannitol and trehalose-specific PTS system, and it also has a certain impact on the transport of several carbohydrates, including glucose, galactose, lactose, and maltose [Citation32]. The deletion of the rcrR gene in S. mutans decreases the ability to form the biofilm and affects the expression of its multiple genes related to sugar transportation [Citation32].

StsR is the transcription factors of GntR family that specifically binds to the predicted promoter sequences in the mannitol-specific PTS transporter, the mannose-specific PTS transporter, maltose ABC transporter and multiple sugar-binding ABC transporter in S. mutans [Citation33]. When stsR is knocked out, the amount of biofilm and the content of exopolysaccharides of S. mutans decreases significantly at the early stage [Citation33].

AhrC is a transcription factor of the arginine repressor (ArgR) family in S. mutans. The results of electrophoretic mobility shift assay (EMSA) and DNase I footprinting showed that AhrC specifically binds to the argC promoter region [Citation34]. AhrC also binds to the promoter regions of argG, Smu.815, Smu.1904, and Smu.940c to regulate the downstream gene expression [Citation34]. In addition, the S. mutans strain overexpressing ahrC has a significant decrease in the formation of biofilm and content of water-insoluble exopolysaccharides [Citation34]. These results suggest that AhrC negatively regulates arginine biosynthesis and biofilm formation in S. mutans.

FruR is a deoxyribonucleoside repressor (deoR)-type regulator that co-transcribes a single EIIABCFru permease with FruI, and a putative 1-phosphofructokinase (1-PFK) with FruK. The loss of FruR results in a significantly reduction of S. mutans biofilm [Citation35].

PdxR is encoded by smu.864, a member of the GntR superfamily of regulatory proteins. When S. mutans is grown in defined medium, pdxR deficiency affects its growth of and significantly reduces biofilm formation [Citation36].

A total of 130 transcription factors were predicted in S. mutans (https://mistdb.com/). However, the functions of several transcription factors remain unknown. Therefore, further research is necessary to investigate the relationships between the uncharacterized transcription factors and biofilm formation.

Two-component systems

Bacterial two-component systems are regulatory circuits of signal transduction based typically on a membrane bound sensor histidine kinase (HK) and a cytoplasmic response regulator (RR) that is activated through a histidine to aspartate phosphorelay reaction [Citation37]. At present, 14 two-component systems and one orphan regulator (CovR) have been found in S. mutans [Citation38]. Among them, VicR positively regulates the expression of gtfs by binding to the promoter region of gtfB and gtfC [Citation39,Citation40]. CovR (also known as GcrR) is a negative regulator that inhibits the expression of gtfB and gtfC by directly binding to the promoter region [Citation41].

Zhang et al. constructed a series of mutant strains (including vicR+covR+, vicR and covR overexpression; vicR+cov-, vicR overexpression and covR deficiency; ASvicRcovR+, vicR low-expression and covR overexpression; ASvicRcovR−, vicR low-expression and covR deficient) to explore the synergy between VicR and CovR in S. mutans in the regulation of the sucrose-selective exopolysaccharides metabolism, and they found that gtfB/gtfC expression is mainly regulated by covR regardless of the vicR gene expression, as revealed by the analysis of the biofilm content of these several strains [Citation42]. These results suggest that CovR may play a dominant role in the interaction between CovR and VicR affecting the expression of gtfs.

VicK is a histidine protein kinase that monitors and transmits chemical signals to the downstream regulatory proteins such as VicR and CovR in S. mutans [Citation43]. Subsequently, VicR or CovR regulate biofilm-associated genes at a transcriptional level, such as gtfB/C, ftf and gbpB [Citation44]. A study found that the deletion of the vicK gene in S. mutans suppresses biofilm formation as well as exopolysaccharides production, and the expression of genes related to exopolysaccharides synthesis are down-regulated, with the exception of gtfB [Citation45]. In addition, the expression of vicX and covR genes are decreased after vicK knockout. These results indicate that VicK regulates biofilm formation by affecting CovR and VicR.

Regulation of S. mutans biofilm formation at the post-transcriptional level

Small RNAs (sRNAs) in bacteria are typically 50–400 nucleotides in length, playing a key role in regulating gene expression [Citation46,Citation47]. In the canonical pathway regulating sRNA-mediated gene expression, sRNAs, often along with the chaperone protein Hfq, positively or negatively regulate the target mRNA expression in response to environmental changes through an incomplete Watson–Crick base pairing [Citation48]. Some studies discovered the presence of many sRNAs with regulatory functions in S. mutans [Citation49–51]. Yin et al. detected a potential sRNA regulation pathway in S. mutans by bioinformatics technology [Citation52]. The results showed that sRNA0426 has a strong positive relationship with the dynamic biofilm formation in S. mutans. Furthermore, the expression of gtfB and gtfC mRNAs are positively correlated with sRNA0426 expression. Liu et al. evaluated the expression of sRNAs and target genes (gtfB, gtfC, and spaP) related to virulence, and they found that the target mRNA of srn92530 is gtfB, while the target mRNA of srn92532 and srn133489 is gtfB and gtfC, respectively [Citation53].

The rnc gene is considered the coding gene of ribonuclease III (RNase III), which can promote mRNA maturation [Citation54]. Mao et al. observed that the rnc gene is located upstream of the vicRKX tricistronic operon and the downstream vic locus is repressed by rnc at the mRNA level in S. mutans. Three putative microRNA-size small RNAs (msRNAs), such as msRNA1701, msRNA3405, and msRNA1657, are negatively correlated with vicRKX but positively correlated with rnc, as revealed by deep sequencing and bioinformatics analysis [Citation55]. Moreover, they found that the knockout rnc gene reduces the ability of biofilm formation of S. mutans [Citation56].

Antisense RNA (AS RNA) refers to an RNA that inhibits the expression of related genes after complementing with mRNA [Citation57]. Lei et al. detected an AS RNA named ASvicR upstream of the rnc gene in S. mutans, whose overexpression inhibits the transcription of vicR, gtfB, gtfC and gtfD, resulting in the decrease of the exopolysaccharides synthesis and biofilm formation [Citation58,Citation59]. Moreover, msRNA1657, a microRNA-size small RNA, binds to the 5′-UTR region of the vicR gene [Citation59,Citation60]. This suggests that ASvicR and msRNA1657 regulate gtfs transcription through the vicR gene. At present, the detailed mechanisms of small RNA affecting the biofilm formation of S. mutans is still not well-known.

Regulation of S. mutans biofilm formation at the post-translation level

Protein translational modifications (PTMs) increase the functional diversity of proteome through the covalent addition of functional groups or proteins, the proteolytic cleavage of regulatory subunits or the degradation of the whole protein [Citation61,Citation62]. PTM has important functions in regulating the physiological activities of S. mutans. For example, glutathionylation on the Cys41 residue of Tlp is crucial to protect S. mutans from oxidative stress and in the competition with S. sanguinis and S. gordonii [Citation63]. Lei et al. identified 973 acetylation sites in 445 proteins of S. mutans, and among them, 617 acetylation sites in 302 proteins were quantified, revealing that the activity of GtfB, GtfC, and GtfD is regulated by acetylation [Citation64]. A recent study found that the GCN5-related N-acetyltransferases (GNAT) family member ActG catalyzes the acetylation of GtfB and GtfC in S. mutans, subsequently inhibiting the synthesis of water-insoluble exopolysaccharides and biofilm formation [Citation65].

Li et al. performed a global protein lysine malonylation (Kmal) analysis of S. mutans and identified a total of 392 malonyllysine sites in 159 proteins [Citation66]. Among them, proteins (WapA, SpaP, GtfC, Ftf, GbpB) closely related to S. mutans biofilm formation were modified with Kmal, suggesting that Kmal is related to exopolysaccharides synthesis and biofilm formation [Citation66].

Regulation of S. mutans biofilm formation by other mechanisms

CRISPR-associated proteins

CRISPR clusters are DNA repeats family widely present in the genome of bacteria and archaea [Citation67]. A CRISPR locus consists of a CRISPR array bordered by various cas genes, which are composed of short direct repeats separated by short variable DNA sequences (called spacers) [Citation68,Citation69]. Two CRISPR-Cas systems exist in S. mutans: the CRISPR1 system (type II-A) and the CRISPR2 system (type I-C) [Citation70].

A study found that the expression of two CRISPR/Cas locus sites in S. mutans is differentially regulated in cdaA knockout strains, which show an abnormal biofilm phenotype [Citation23]. Zhang et al. found that S. mutans clinical strains with both CRIPSR systems have a stronger ability of biofilm formation [Citation70]. Later, Tang et al. observed that the knockout of cas3 reduces the ability of S. mutans to produce exopolysaccharides and form biofilm by decreasing the expression of gtfB and gtfC [Citation71]. This result suggests that the CRISPR-Cas systems are involved in the regulation of gtfs expression and biofilm formation in S. mutans.

PTS system

The PTS system consists of the nonspecific enzyme I (EI), histidine-containing the phosphocarrier protein (HPr), and sugar-specific enzyme II (EII) complexes. The EI and HPr are involved in the transport of all PTS sugars: HPr encodes ptsH and EI encodes ptsI in S. mutans [Citation72]. The EII complex usually consists of three parts: EIIA, EIIB and EIIC, where EIIA and EIIB are located in the cytoplasm and EIIC is an integral membrane protein (mannose-type PTS system containing EIID) [Citation73].

EIIMan is the most physiologically important PTS complex in S. mutans, mainly involved in the transport of glucose, mannose, galactose, glucosamine (GlcN) and N-acetylglucosamine (GlcNAc) [Citation74,Citation75]. This complex is made up of four domains expressed as three polypeptides in a single operon: EIIAB (manL), EIIC (manM), and EIID (manN). Previous studies showed that the deletion of the manL gene relieved CCR of several carbohydrate catabolic operons, including the cel operon encoding a phospho-β-glucosidase (CelA), a cellobiose-PTS EII complex [Citation76,Citation77], the lac operon encoding EIILac (lacFE) and proteins required for the utilization of both lactose and galactose [Citation74]. Abranches et al. found that the reduced expression of the gtfBC promoter in the EIIABMan mutant strain is associated to the involvement of ManL in the regulation of gtfBC expression [Citation78].

Quorum-sensing signaling system

Quorum sensing is a mechanism of bacterial intercellular communication for the regulation of gene expression in response to the density of the population. It is based on the production, detection and response of extracellular signaling molecules known as autoinducers [Citation79]. The LuxS system mediates the interspecific interactions in multi-species communities, and it is one of the widely studied quorum-sensing systems in S. mutans [Citation38].

The luxS gene is involved in many cellular processes, and it is present in a wide range of bacteria [Citation80]. LuxS catalyzes the formation of the autoinducer AI−2 (a furanosyl borate diester) in the methyl cycle [Citation81]. A study found that the knockout of luxS in S. mutans suppresses gtfD expression [Citation82]. In addition, its knockdown affects S. mutans ABC transporters and carbohydrate transport, transformation, and metabolism through the EII subunits and enzymes to influence virulence-associated traits, as revealed by a transcriptome analysis [Citation81]. Thus, the quorum-sensing signaling system in S. mutans is involved in the formation of biofilms.

Enzymes

Dextranase (Dex) is a glucanase involved in the degradation of water-soluble glucans, and includes DexA and DexB [Citation83]. DexA breaks the α−1,6-linkage of the extracellular glucans to produce oligosaccharides, which are degraded into monosaccharides by DexB glucosidase after entering the cells [Citation84–86]. Yang et al. revealed that dexA knockout results in an increased transcription of genes related to exopolysaccharides synthesis, including gtfB, gtfD and ftf [Citation87]. In addition, the biofilms of the strains overexpressing dexA lack exopolysaccharides matrix and are unable to aggregate into dense and continuous microcolonies [Citation87]. These results indicate an important role of the dexA gene in the formation of S. mutans biofilm.

YidC proteins are membrane-localized chaperone insertases that are universally conserved in all bacteria [Citation88]. The genome of S. mutans contains two copies of genes encoding the YidC homologs yidC1 and yidC2, and their protein sequences are 27% identical and 48% similar. The deletion of yidC1 or yidC2 results in a reduced synthesis of insoluble glucan at the early and mid-exponential growth phases [Citation89]. In particular, the deletion of yidC2 resulted in a significant reduction in biofilm biomass, evident defects in the spatial organization of the extracellular polymer matrix, and alteration in the three-dimensional biofilm structure [Citation89]. The defective biofilm contains smaller bacterial clusters with higher cell density and less surrounding exopolysaccharides than the biofilm in the wild type [Citation89].

The copYAZ is a copper-transport operon in S. mutans that encodes three proteins: a copper P-type ATPases (CopA), a copper-responsive repressor (CopY), and a copper chaperone (CopZ) [Citation90]. CopY allows transcription upon copper transfer from direct binding with CopZ [Citation90,Citation91]. CopA is a small protein that tightly binds to copper and delivers it to CopY to positively regulate copYAZ [Citation90–92]. CopZ exports the excess copper ions to the outside of the cell, causing the cell to become resistant to copper [Citation90,Citation91]. Garcia et al. found that glycosyltransferase secretion and biofilm formation are significantly reduced after copZ knockdown in S. mutans [Citation93]. This result suggests that CopZ is a global regulator of biofilms. Further characterization of CopZ may lead to the identification of novel pathways in biofilm formation and characterization.

EzrA is an integral membrane protein consisting of an N-terminal trans-membrane (TM) spanning helix followed by an 60 kDa cytoplasmic domain [Citation94]. It plays a key role in the regulation of cell division and in the maintenance of cell size and shape [Citation95–97]. Xiang et al. investigated the role of EzrA in S. mutans by constructing ezrA in-frame deletion mutants. The results showed that the ezrA mutant grows slower that the wild type, with a round cell shape characterized by a shortened length and extended width [Citation98]. Single-species cariogenic biofilm model revealed that the deletion of ezrA results in a defective biofilm formation with less exopolysaccharides and altered three-dimensional biofilm architecture [Citation98].

Amyloid is a fibrous cross-β sheet quaternary structure consisting of ordered peptides or polymerized proteins [Citation99]. S. mutans is an amyloid-forming organism and amyloidogenesis contributes to biofilm formation in this bacterium [Citation100]. Three proteins are present in S. mutans that form amyloidogenic fibrils, including amyloidogenic adhesin SpaP, wall-associated protein A (WapA) and the secreted protein Smu.63c [Citation101]. When S. mutans grows in biofilm medium containing glucose as the sole carbon source, the strains with spaP and wapA deletion produce significantly less biofilm than the wild type strain, while the deletion of Smu.63c results in a slight increase in biofilm production [Citation101]. The reduction of biofilm is more significant in the ΔspaP/ΔwapA double-deletion strain and the ΔspaPsmu.63cwapA triple-deletion strain [Citation101]. ΔSmu.63cwapA double-deletion strain and ΔSmu.63cspaP double-deletion strain rescued the spaP and wapA deletion phenotypes, resulting in biofilm production comparable to that of the wild-type strain [Citation101]. These results suggest that SpaP and WapA positively regulate the biofilm formation in S. mutans, while Smu.63c exerts a negative regulatory role.

The efflux pump is a protein localized on the cell membrane that allows the maintenance of the homeostasis of the internal environment in the microorganisms by excreting toxic substances, such as antimicrobial drugs, metabolites and signal molecules of the quorum sensing [Citation102]. LmrB is a putative efflux pump in S. mutans, and its inactivation results in the structural changes of the biofilm, increased production of exopolysaccharides and increased transcription of exopolysaccharides-related genes [Citation103]. A number of genes involved in sugar uptake and metabolism are up-regulated, including sugar metabolism associated glg operons and msmREFGK transporter, as demonstrated by the transcriptome analysis [Citation104].

The smu.833 gene encodes a hypothetical glycosyltransferase highly conserved among the group of mutans streptococci but absent in other oral commensals [Citation105]. The protein encoded by smu.833 has two transmembrane helices and is highly homologous to proteins involved in cell wall biogenesis and the response to cellular stress [Citation105]. Rainey et al. found that smu.833 regulates the dynamic interaction between glucan and extracellular DNA (eDNA) of S. mutans [Citation105]. Moreover, the loss of smu.833 resulted in a significant decrease in the expression of Gtfs and a concurrent reduction in glucan matrix [Citation105].

The streptococcal pleiotropic regulator of virulence (SprV) is a small hypothetical protein that is encoded by smu.2137 in S. mutans. The deletion of SprV leads to an impaired biofilm formation and various virulence-related functions in S. mutans [Citation106]. Transcriptome sequencing revealed that gtfB, gbpA, gbpB, and gbpC, which are genes related to biofilm formation, are downregulated by 2.6-, 2.4-, 1.7-, and 2.6-fold, respectively [Citation106].

Recombinase A (RecA) is related to the SOS-response in S. mutans. The RecA-deficient mutant strain possesses a lower acid tolerance and produces a biofilm with a lower density than the wild type [Citation107]. Glutamate racemase (MurI) is an essential enzyme for the biosynthesis of peptidoglycan. murI deficiency in S. mutans weakens the ability to form the biofilm and virulence factors. The expression of comE, comD, gtfB and gtfC, genes related to biofilm formation are down-regulated 8-, 43-, 85- and 298-fold, as revealed by qRT-PCR [Citation108].

Small-molecule metabolites

Small molecules are often encoded by biosynthetic gene clusters (BGCs) and are the primary means of communication in the microbial world. They are involved in various microbial–microbial and host-microbial interactions, including antimicrobial activity, bacterial signaling, immunomodulation, biofilm formation, host colonization, nutrient clearance, and stress protection [Citation109]. The genomic analysis of S. mutans isolates revealed that it collectively harbors a plethora of BGCs for polyketide/non-ribosomal peptide biosynthesis [Citation110].

The polyketide/nonribosomal peptide biosynthetic gene cluster (mufD, mufE, mufF, mufG) was first identified in S. mutans strains isolated from dental plaque [Citation111]. Zhong et al. found that the muf gene cluster synthesizes the small-molecule mutanofactin−697, as demonstrared by liquid chromatography – high-resolution mass spectrometry (LC – HRMS) [Citation111]. Further mode-of-action studies showed that mutanofactin−697 binds to S. mutans and extracellular DNA, increases bacterial hydrophobicity and promotes bacterial adhesion and subsequent biofilm formation () [Citation111].

Table 1. Summary of genes regulating biofilm formation in S. mutans.

The strategies for biofilm inhibition and caries prevention

Although caries is a multi-microbial disease, selective targeting of S. mutans has been considered to be an appropriate approach for the prevention of dental caries [Citation112]. This is mainly because the synthesis of exopolysaccharides from sucrose by S. mutans is important to the development of cariogenic biofilms. Thus, targeted inhibition of S. mutans biofilm may be a viable approach to impede the progression of caries without disrupting the oral microbiome associated with health.

Gtfs secreted by S. mutans can synthesize exopolysaccharides using sucrose [Citation113,Citation114]. Therefore, targeting Gtfs could selectively prevent the synthesis of exopolysaccharides, the formation of biofilm and impair the S. mutans virulence, which do not threaten the microecological balance in the oral cavity. Screening for compounds that inhibit Gtfs could have a significant effect on biofilm inhibition. Fluoride can inhibit the production and secretion of S. mutans Gtfs [Citation115]. Some sucrose structural analogs (including 6-deoxysucrose, 6-thiosucrose, 4,6-dideoxysucrose, sucralose, 4-deoxysucrose, 4-chloro−4-deoxygalactosucrose, 6,6’-dithiodisucrose) and natural products (including Tris, green mate, roasted mate, 7-epiclusianone, cranberry juice, quercetin, polyphenols) could also inhibit Gtfs activities [Citation116–123]. Because the specificity of these compounds is largely unknown, their access to clinical applications remains limited.

The upstream regulatory genes, including TCS and quorum sensing system, regulate S. mutans biofilm formation by affecting the promoter activity of gtfs and subsequent expression of Gtfs. Therefore, these upstream regulatory systems may be potential targets, which can be used to develop therapeutic agents against dental caries. A number of inhibitors targeting upstream regulatory pathways have been identified. WalKHK, a TCS inhibitor from B. subtilis, inhibits the in vitro autophosphorylation of VicK in S. mutans [Citation124]. The cell extract from Tenacibaculum sp. 20J interfered specifically with the AI−2 quorum sensing system and reduced biofilm formation of S. mutans [Citation125]. Although the expression of gtfs is well regulated, the exact mechanism of regulation is still not fully understood. If these regulatory genes also affect the expression of other genes (e.g. mutacins), then the strategies of targeting upstream regulatory genes may affect the oral microecological homeostasis.

Second messenger c-di-AMP and Ap4A regulate S. mutans biofilms formation. Inhibitors targeting synthetase/hydrolase or receptor proteins can affect the intracellular levels or signaling transduction of second messengers and inhibit biofilm formation. Currently, most c-di-AMP inhibitors target PDE and DisA [Citation126]. For example, three inhibitors of Bacillus subtilis DisA, bromophenol-TH, suramin, and the tea polyphenol theaflavin digallate, were identified [Citation126]. Stress alarmone ppGpp could competitively inhibit B. subtilis GdpP (PDE) [Citation127]. However, inhibitors targeting S. mutans DAC/PDE still need further investigation. The mechanisms of action of these inhibitors and whether there might be an impact on the other flora in the oral cavity are still unclear.

Conclusions

S. mutans is the core microorganism causing human dental caries due to its high ability of biofilm formation and cariogenicity. The regulatory network of biofilm formation mainly includes upstream signals composed of c-di-AMP and Ap4A, regulating gtfs gene expression by CRISPR-Cas, transcription factor, two-component systems, small RNA, PTS systems, and modulating glucosyltransferases activity by acetylation, malonylation. The specific inhibition of S. mutans biofilm modulates the oral microecological balance. Therefore, targeting the systems that regulate the biofilm formation might provide potential strategies for the ecological prevention and treatment of caries.

Author contributions

T. Zheng, X. Zhou, J. Zeng, Y. Li contributed to conception, design, and drafted and critically revised the manuscript; M. Jing, T. Gong, J. Yan, X. Wang, M. Xu contributed to conception, analysis, interpretation and critically revised the manuscript. All authors gave final approval and agreed to be accountable for all aspects of the work.

Acknowledgments

This work was supported by grants from National Natural Science Foundation of China [32170046], the Sichuan Science and Technology Program [2022YFH0048], and State Key Laboratory of Oral Disease [2022KXK0403].

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by grants from National Natural Science Foundation of China [32170046], the Sichuan Science and Technology Program [2022YFH0048], and State Key Laboratory of Oral Disease [2022KXK0403].

References

  • Pitts NB, Zero DT, Marsh PD, et al. Dental caries, nature reviews. Disease Primers. 2017;3(1):17030. doi: 10.1038/nrdp.2017.30
  • Sabharwal A, Stellrecht E, Scannapieco FA. Associations between dental caries and systemic diseases: a scoping review. BMC Oral Health. 2021;21(1). doi: 10.1186/s12903-021-01803-w
  • Vos T, Lim SS, Abbafati C. Global burden of 369 diseases and injuries in 204 countries and territories, 1990–2019: a systematic analysis for the global burden of disease study 2019. Lancet. 2020;396(10258):1204–13. doi: 10.1016/S0140-6736(20)30925-9
  • Wen PYF, Chen MX, Zhong YJ, et al. Global burden and inequality of dental caries, 1990 to 2019. J Dent Res. 2022;101(4):392–399. doi: 10.1177/00220345211056247
  • Peres MA, Macpherson LMD, Weyant RJ, et al. Oral diseases: a global public health challenge. Lancet. 2019;394(10194):249–260. doi: 10.1016/S0140-6736(19)31146-8
  • Lin Y, Chen J, Zhou X, et al. Inhibition of Streptococcus mutans biofilm formation by strategies targeting the metabolism of exopolysaccharides. Crit Rev Microbiol. 2021;47(5):667–677. doi: 10.1080/1040841X.2021.1915959
  • Lemos JA, Palmer SR, Zeng L, et al. The biology of Streptococcus mutans. Microbiol Spectr. 2019;7(1). doi: 10.1128/microbiolspec.GPP3-0051-2018
  • Mieher JL, Larson MR, Schormann N, et al. Glucan binding protein C of Streptococcus mutans mediates both sucrose-independent and sucrose-dependent adherence. Infect Immun. 2018;86(7). doi: 10.1128/IAI.00146-18
  • Krzysciak W, Jurczak A, Koscielniak D, et al. The virulence of Streptococcus mutans and the ability to form biofilms. Eur J Clin Microbiol Infect Dis. 2014;33(4):499–515. doi: 10.1007/s10096-013-1993-7
  • Cross BW, Ruhl S. Glycan recognition at the saliva - oral microbiome interface. Cell Immunol. 2018;333:19–33. doi: 10.1016/j.cellimm.2018.08.008
  • Brady LJ, Maddocks SE, Larson MR, et al. The changing faces of Streptococcus antigen I/II polypeptide family adhesins. Mol Microbiol. 2010;77(2):276–286. doi: 10.1111/j.1365-2958.2010.07212.x
  • Abranches J, Zeng L, Kajfasz JK, et al. Biology of oral streptococci, Microbiol Spectr. 2018;6(5). doi: 10.1128/microbiolspec.GPP3-0042-2018
  • Zhang Q, Ma Q, Wang Y, et al. Molecular mechanisms of inhibiting glucosyltransferases for biofilm formation in Streptococcus mutans. Int J Oral Sci. 2021;13(1):30. doi: 10.1038/s41368-021-00137-1
  • Klein MI, Hwang G, Santos PH, et al. Streptococcus mutans-derived extracellular matrix in cariogenic oral biofilms. Front Cell Infect Microbiol. 2015;5:10. doi: 10.3389/fcimb.2015.00010
  • Karygianni L, Ren Z, Koo H, et al. Biofilm matrixome: extracellular components in structured microbial communities. Trends Microbiol. 2020;28(8):668–681. doi: 10.1016/j.tim.2020.03.016
  • Ellepola K, Truong T, Liu Y, et al. Multi-omics analyses reveal synergistic carbohydrate metabolism in Streptococcus mutans-Candida albicans mixed-species biofilms. Infect Immun. 2019;87(10). doi: 10.1128/IAI.00339-19
  • Banerjee P, Sahoo PK, Sheenu AA, et al. Molecular and structural facets of c-di-GMP signalling associated with biofilm formation in Pseudomonas aeruginosa. Mol Aspects Med. 2021;81:101001. doi: 10.1016/j.mam.2021.101001
  • Commichau FM, Dickmanns A, Gundlach J, et al. A jack of all trades: the multiple roles of the unique essential second messenger cyclic di-AMP. Mol Microbiol. 2015;97(2):189–204. doi: 10.1111/mmi.13026
  • Bai Y, Yang J, Zhou X, et al. Mycobacterium tuberculosis Rv3586 (DacA) is a diadenylate cyclase that converts ATP or ADP into c-di-AMP. PLoS ONE. 2012;7(4):e35206. doi: 10.1371/journal.pone.0035206
  • Xiong ZQ, Fan YZ, Song X, et al. The second messenger c-di-AMP mediates bacterial exopolysaccharide biosynthesis: a review. Mol Biol Rep. 2020;47(11):9149–9157. doi: 10.1007/s11033-020-05930-5
  • Ye M, Zhang JJ, Fang X, et al. DhhP, a cyclic di-AMP phosphodiesterase of Borrelia burgdorferi, is essential for cell growth and virulence. Infect Immun. 2014;82(5):1840–1849. doi: 10.1128/IAI.00030-14
  • Huynh TN, Luo S, Pensinger D, et al. An HD-domain phosphodiesterase mediates cooperative hydrolysis of c-di-AMP to affect bacterial growth and virulence. Proc Natl Acad Sci USA. 2015;112(7):E747–56. doi: 10.1073/pnas.1416485112
  • Cheng X, Zheng X, Zhou X, et al. Regulation of oxidative response and extracellular polysaccharide synthesis by a diadenylate cyclase in Streptococcus mutans. Environ Microbiol. 2016;18(3):904–922. doi: 10.1111/1462-2920.13123
  • Peng X, Michalek S, Wu H. Effects of diadenylate cyclase deficiency on synthesis of extracellular polysaccharide matrix of Streptococcus mutans revisit. Environ Microbiol. 2016;18(11):3612–3619. doi: 10.1111/1462-2920.13440
  • Peng X, Zhang Y, Bai G, et al. Cyclic di-AMP mediates biofilm formation. Mol Microbiol. 2016;99(5):945–959. doi: 10.1111/mmi.13277
  • Ji X, Zou J, Peng HB, et al. Alarmone Ap4A is elevated by aminoglycoside antibiotics and enhances their bactericidal activity. Proc Natl Acad Sci USA. 2019;116(19):9578–9585. doi: 10.1073/pnas.1822026116
  • Randerath K, Janeway CM, Stephenson ML, et al. Isolation and characterization of dinucleoside tetra- and tri-phosphates formed in the presence of lysyl-Srna synthetase. Biochem Biophys Res Commun. 1966;24(1):98–105. doi: 10.1016/0006-291X(66)90416-5
  • Ofir-Birin Y, Fang PF, Bennett SP, et al. Structural switch of Lysyl-Trna synthetase between translation and transcription. Mol Molecular Cell. 2013;49(1):30–42. doi: 10.1016/j.molcel.2012.10.010
  • Ferguson F, McLennan AG, Urbaniak MD, et al. Copeland, Re-evaluation of diadenosine tetraphosphate (Ap4a) from a stress metabolite to bona fide secondary messenger. Front Mol Biosci. 2020;7:606807. doi: 10.3389/fmolb.2020.606807
  • Zheng T, Jing M, Gong T, et al. Deletion of the yqeK gene leads to the accumulation of Ap4A and reduced biofilm formation in Streptococcus mutans. Mol Oral Microbiol. 2021;37(1): 9–21. doi: 10.1111/omi.12356
  • Chen J, Zhang A, Xiang Z, et al. EpsR negatively regulates Streptococcus mutans exopolysaccharide synthesis. J Dent Res. 2021;100(9):968–976. doi: 10.1177/00220345211000668
  • Gong T, He X, Chen J, et al. Transcriptional profiling reveals the importance of RcrR in the regulation of multiple sugar transportation and biofilm formation in Streptococcus mutans. mSystems. 2021;6(4):e0078821. doi: 10.1128/mSystems.00788-21
  • Li Z, Xiang Z, Zeng J, et al. A GntR family transcription factor in Streptococcus mutans regulates biofilm formation and expression of multiple sugar transporter genes. Front Microbiol. 2018;9:3224. doi: 10.3389/fmicb.2018.03224
  • Jing ML, Zheng T, Gong T, et al. AhrC negatively regulates Streptococcus mutans arginine biosynthesis. Microbiol Spectr. 2022;10(4): e0072122. doi: 10.1128/spectrum.00721-22
  • Zeng L, Chakraborty B, Farivar T, et al. Coordinated regulation of the EIIMan an fruRKI operons of Streptococcus mutans by global and fructose-specific pathways. Environ Microbiol. 2017;83(21). doi: 10.1128/AEM.01403-17
  • Liao S, Bitoun JP, Nguyen AH, et al. Deficiency of PdxR in Streptococcus mutans affects vitamin B6 metabolism, acid tolerance response and biofilm formation. Mol Oral Microbiol. 2015;30(4):255–268. doi: 10.1111/omi.12090
  • Padilla-Vaca F, Mondragon-Jaimes V, Franco B. General aspects of two-component regulatory circuits in bacteria: domains, signals and roles. Curr Protein Pept Sci. 2017;18(10):990–1004. doi: 10.2174/1389203717666160809154809
  • Smith EG, Spatafora GA. Gene regulation in S. mutans: complex control in a complex environment. J Dent Res. 2012;91(2):133–141. doi: 10.1177/0022034511415415
  • Senadheera MD, Guggenheim B, Spatafora GA, et al. A VicRK signal transduction system in Streptococcus mutans affects gtfBCD, gbpB, and ftf expression, biofilm formation, and genetic competence development. J Bacteriol. 2005;187(12):4064–4076. doi: 10.1128/JB.187.12.4064-4076.2005
  • Lei L, Long L, Yang X, et al. The VicRK two-component system regulates Streptococcus mutans virulence. Curr Issues Mol Biol. 2019;32:167–200. doi: 10.21775/cimb.032.167
  • Biswas S, Biswas I. Regulation of the glucosyltransferase (gtfBC) operon by CovR in Streptococcus mutans. J Bacteriol. 2006;188(3):988–998. doi: 10.1128/JB.188.3.988-998.2006
  • Zhang HY, Xia MY, Zhang B, et al. Sucrose selectively regulates Streptococcus mutans polysaccharide by GcrR. Environ Microbiol. 2022;24(3):1395–1410. doi: 10.1111/1462-2920.15887
  • Downey JS, Mashburn-Warren L, Ayala EA, et al. In vitro manganese-dependent cross-talk between Streptococcus mutans VicK and GcrR: implications for overlapping stress response pathways. PLoS ONE. 2014;9(12):e115975. doi: 10.1371/journal.pone.0115975
  • Stipp RN, Boisvert H, Smith DJ, et al. CovR and VicRK regulate cell surface biogenesis genes required for biofilm formation in Streptococcus mutans. PLoS ONE. 2013;8(3):e58271. doi: 10.1371/journal.pone.0058271
  • Deng Y, Yang Y, Zhang B, et al. The vicK gene of Streptococcus mutans mediates its cariogenicity via exopolysaccharides metabolism. Int J Oral Sci. 2021;13(1):45. doi: 10.1038/s41368-021-00149-x
  • Storz G, Vogel J, Wassarman KM. Regulation by small RNAs in bacteria: expanding frontiers. Mol Molecular Cell. 2011;43(6):880–891. doi: 10.1016/j.molcel.2011.08.022
  • Wagner EGH, Romby P. Small RNAs in bacteria and archaea: who they are, what they do, and how they do it. Adv Genet. 2015;90(90):133–208. doi: 10.1016/bs.adgen.2015.05.001
  • Reyer MA, Chennakesavalu S, Heideman EM, et al. Kinetic modeling reveals additional regulation at co-transcriptional level by post-transcriptional sRNA regulators. Cell Rep. 2021;36(13):109764. doi: 10.1016/j.celrep.2021.109764
  • Lee HJ, Hong SH. Analysis of microRNA-size, small RNAs in Streptococcus mutans by deep sequencing. FEMS Microbiol Lett. 2012;326(2):131–136. doi: 10.1111/j.1574-6968.2011.02441.x
  • Zhu WH, Liu SS, Liu J, et al. High-throughput sequencing identification and characterization of potentially adhesion-related small RNAs in Streptococcus mutans. J Med Microbiol. 2018;67(5):641–651. doi: 10.1099/jmm.0.000718
  • Krieger MC, Merritt J, Raghavan R, et al. Genome-wide identification of novel sRnas in Streptococcus mutans. J Bacteriol. 2022;204(4). doi: 10.1128/jb.00577-21
  • Yin L, Zhu W, Chen D, et al. Small noncoding RNA sRNA0426 is involved in regulating biofilm formation in Streptococcus mutans. Microbiologyopen. 2020;9(9):e1096. doi: 10.1002/mbo3.1096
  • Liu SS, Zhu WH, Zhi QH, et al. Analysis of sucrose-induced small RNAs in Streptococcus mutans in the presence of different sucrose concentrations. Appl Microbiol Biotechnol. 2017;101(14):5739–5748. doi: 10.1007/s00253-017-8346-x
  • Viegas SC, Silva IJ, Saramago M, et al. Regulation of the small regulatory RNA MicA by ribonuclease III: a target-dependent pathway. Nucleic Acids Res. 2011;39(7):2918–2930. doi: 10.1093/nar/gkq1239
  • Mao MY, Yang YM, Li KZ, et al. The rnc gene promotes exopolysaccharide synthesis and represses the vicRKX gene expressions via microRNA-Size small RNAs in Streptococcus mutans. Front Microbiol. 2016;7:678. doi: 10.3389/fmicb.2016.00687
  • Mao MY, Li M, Lei L, et al. The regulator gene rnc is closely involved in biofilm formation in Streptococcus mutans. Caries Res. 2018;52(5):347–358. doi: 10.1159/000486431
  • Kulkarni JA, Witzigmann D, Thomson SB, et al. The current landscape of nucleic acid therapeutics. Nat Nanotechnol. 2021;16(6):630–643. doi: 10.1038/s41565-021-00898-0
  • Lei L, Stipp RN, Chen T, et al. Activity of Streptococcus mutans VicR is modulated by antisense RNA. J Dent Res. 2018;97(13):1477–1484. doi: 10.1177/0022034518781765
  • Lei L, Zhang B, Mao M, et al. Carbohydrate metabolism regulated by antisense vicR RNA in cariogenicity. J Dent Res. 2020;99(2):204–213. doi: 10.1177/0022034519890570
  • Lei L, Yang Y, Wu S, et al. Mechanisms by which small RNAs affect bacterial activity. J Dent Res. 2019;98(12):1315–1323. doi: 10.1177/0022034519876898
  • Moon SP, Balana AT, Pratt MR. Consequences of post-translational modifications on amyloid proteins as revealed by protein semisynthesis. Curr Opin Chem Biol. 2021;64:76–89. doi: 10.1016/j.cbpa.2021.05.007
  • Ma QZ, Zhang Q, Chen Y, et al. Post-translational modifications in oral bacteria and their functional impact. Front Microbiol. 2021;12. doi: 10.3389/fmicb.2021.784923
  • Li ZY, Zhang CZ, Li C, et al. S-glutathionylation proteome profiling reveals a crucial role of a thioredoxin-like protein in interspecies competition and cariogenecity of Streptococcus mutans. PLOS Pathog. 2020;16(7):e1008774. doi: 10.1371/journal.ppat.1008774
  • Lei L, Zeng JM, Wang LY, et al. Quantitative acetylome analysis reveals involvement of glucosyltransferase acetylation in Streptococcus mutans biofilm formation. Env Microbiol Rep. 2021;13(2):86–97. doi: 10.1111/1758-2229.12907
  • Ma QZ, Pan YY, Chen Y, et al. Acetylation of glucosyltransferases regulates Streptococcus mutans biofilm formation and virulence. PLOS Pathog. 2021;17(12):e1010134. doi: 10.1371/journal.ppat.1010134
  • Li Z, Wu Q, Zhang Y, et al. Systematic analysis of lysine malonylation in Streptococcus mutans. Front Cell Infect Microbiol. 2022;12:1078572. doi: 10.3389/fcimb.2022.1078572
  • Marraffini LA. CRISPR-Cas immunity in prokaryotes. Nature. 2015;526(7571):55–61. doi: 10.1038/nature15386
  • Makarova KS, Wolf YI, Alkhnbashi OS, et al. An updated evolutionary classification of CRISPR-Cas systems. Nat Rev Microbiol. 2015;13(11):722–736. doi: 10.1038/nrmicro3569
  • Gong T, Zeng J, Tang B, et al. CRISPR-Cas systems in oral microbiome: from immune defense to physiological regulation. Mol Oral Microbiol. 2020;35(2):41–48. doi: 10.1111/omi.12279
  • Zhang A, Chen J, Gong T, et al. Deletion of csn2 gene affects acid tolerance and exopolysaccharide synthesis in Streptococcus mutans. Mol Oral Microbiol. 2020;35(5):211–221. doi: 10.1111/omi.12308
  • Tang B, Gong T, Zhou X, et al. Deletion of cas3 gene in Streptococcus mutans affects biofilm formation and increases fluoride sensitivity. Arch Oral Biol. 2019;99:190–197. doi: 10.1016/j.archoralbio.2019.01.016
  • Zeng L, Burne RA. Seryl-phosphorylated HPr regulates CcpA-independent carbon catabolite repression in conjunction with PTS permeases in Streptococcus mutans. Mol Microbiol. 2010;75(5):1145–1158. doi: 10.1111/j.1365-2958.2009.07029.x
  • Jeckelmann JM, Erni B. The mannose phosphotransferase system (Man-PTS) - Mannose transporter and receptor for bacteriocins and bacteriophages. Biochim Biophys Acta Biomembr. 2020;1862(11):183412. doi: 10.1016/j.bbamem.2020.183412
  • Zeng L, Das S, Burne RA. Utilization of lactose and galactose by Streptococcus mutans: transport, toxicity, and carbon catabolite repression. J Bacteriol. 2010;192(9):2434–2444. doi: 10.1128/JB.01624-09
  • Moye ZD, Burne RA, Zeng L, et al. Uptake and metabolism of N-Acetylglucosamine and glucosamine by Streptococcus mutans. Appl Environ Microbiol. 2014;80(16):5053–5067. doi: 10.1128/AEM.00820-14
  • Abranches J, Candella MM, Wen ZZT, et al. Different roles of EIIABMan and EIIGlc in regulation of energy metabolism, biofilm development, and competence in Streptococcus mutans. J Bacteriol. 2006;188(11):3748–3756. doi: 10.1128/JB.00169-06
  • Zeng L, Burne RA. Transcriptional regulation of the cellobiose operon of Streptococcus mutans. J Bacteriol. 2009;191(7):2153–2162. doi: 10.1128/JB.01641-08
  • Abranches J, Chen YYM, Burne RA. Characterization of Streptococcus mutans strains deficient in EIIAB Man of the sugar phosphotransferase system. Appl Environ Microbiol. 2003;69(8):4760–4769. doi: 10.1128/AEM.69.8.4760-4769.2003
  • Mukherjee S, Bassler BL. Bacterial quorum sensing in complex and dynamically changing environments. Nat Rev Microbiol. 2019;17(6):371–382. doi: 10.1038/s41579-019-0186-5
  • Rezzonico F, Duffy B. Lack of genomic evidence of AI-2 receptors suggests a non-quorum sensing role for luxS in most bacteria. BMC Microbiol. 2008;8(1):154. doi: 10.1186/1471-2180-8-154
  • Yuan K, Hou L, Jin Q, et al. 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. doi: 10.1016/j.archoralbio.2021.105137
  • Hu X, Wang Y, Gao L, et al. The impairment of methyl metabolism from luxS mutation of Streptococcus mutans. Front Microbiol. 2018;9:404. doi: 10.3389/fmicb.2018.00404
  • Khalikova E, Susi P, Usanov N, et al. Purification and properties of extracellular dextranase from a Bacillus sp. J Chromatogr B Analyt Technol Biomed Life Sci. 2003;796(2):315–326. doi: 10.1016/j.jchromb.2003.08.037
  • Suzuki N, Kim YM, Fujimoto Z, et al. Structural elucidation of dextran degradation mechanism by Streptococcus mutans dextranase belonging to glycoside hydrolase family 66. J Biol Chem. 2012;287(24):19916–19926. doi: 10.1074/jbc.M112.342444
  • Saburi W, Hondoh H, Unno H, et al. Crystallization and preliminary X-ray analysis of Streptococcus mutans dextran glucosidase. Acta Crystallogr Sect F: Struct Biol Cryst Commun. 2007;63(Pt 9):774–776. doi: 10.1107/S174430910703936X
  • Saburi W, Mori H, Saito S, et al. Structural elements in dextran glucosidase responsible for high specificity to long chain substrate. Biochim Biophys Acta. 2006;1764(4):688–698. doi: 10.1016/j.bbapap.2006.01.012
  • Yang Y, Mao M, Lei L, et al. Regulation of water-soluble glucan synthesis by the Streptococcus mutans dexA gene effects biofilm aggregation and cariogenic pathogenicity. Mol Oral Microbiol. 2019;34(2):51–63. doi: 10.1111/omi.12253
  • Hasona A, Crowley PJ, Levesque CM, et al. Streptococcal viability and diminished stress tolerance in mutants lacking the signal recognition particle pathway or YidC2. Proc Natl Acad Sci USA. 2005;102(48):17466–17471. doi: 10.1073/pnas.0508778102
  • Palmer SR, Ren Z, Hwang G, et al. Streptococcus mutans yidC1 and yidC2 impact cell envelope biogenesis, the biofilm matrix, and biofilm biophysical properties. J Bacteriol. 2019;201(1). doi: 10.1128/JB.00396-18
  • Vats N, Lee SF. Characterization of a copper-transport operon, copYAZ, from Streptococcus mutans. Microbiol (Read). 2001;147(Pt 3):653–662. doi: 10.1099/00221287-147-3-653
  • Cobine P, Wickramasinghe WA, Harrison MD, et al. The Enterococcus hirae copper chaperone CopZ delivers copper(I) to the CopY repressor. FEBS Lett. 1999;445(1):27–30. doi: 10.1016/S0014-5793(99)00091-5
  • Singh K, Senadheera DB, Levesque CM, et al. The copYAZ operon functions in copper efflux, biofilm formation, genetic transformation, and stress tolerance in Streptococcus mutans. J Bacteriol. 2015;197(15):2545–2557. doi: 10.1128/JB.02433-14
  • Garcia SS, Du Q, Wu H. Streptococcus mutans copper chaperone, CopZ, is critical for biofilm formation and competitiveness. Mol Oral Microbiol. 2016;31(6):515–525. doi: 10.1111/omi.12150
  • Levin PA, Kurtser IG, Grossman AD. Identification and characterization of a negative regulator of FtsZ ring formation in Bacillus subtilis. Proc Natl Acad Sci U S A. 1999;96(17):9642–9647. doi: 10.1073/pnas.96.17.9642
  • Jorge AM, Hoiczyk E, Gomes JP, et al. EzrA contributes to the regulation of cell size in Staphylococcus aureus. PLoS ONE. 2011;6(11):e27542. doi: 10.1371/journal.pone.0027542
  • Duman R, Ishikawa S, Celik I, et al. Structural and genetic analyses reveal the protein SepF as a new membrane anchor for the Z ring. Proc Natl Acad Sci USA. 2013;110(48):E4601–E4610. doi: 10.1073/pnas.1313978110
  • Cleverley RM, Barrett JR, Basle A, et al. Structure and function of a spectrin-like regulator of bacterial cytokinesis. Nat Commun. 2014;5(1). doi: 10.1038/ncomms6421
  • Xiang ZT, Li ZB, Ren Z, et al. EzrA, a cell shape regulator contributing to biofilm formation and competitiveness in Streptococcus mutans. Mol Oral Microbiol. 2019;34(5):194–208. doi: 10.1111/omi.12264
  • Nilsson MR. Techniques to study amyloid fibril formation in vitro. Methods. 2004;34(1):151–160. doi: 10.1016/j.ymeth.2004.03.012
  • Oli MW, Otoo HN, Crowley PJ, et al. Functional amyloid formation by Streptococcus mutans. Microbiol-Sgm. 2012;158(12):2903–2916. doi: 10.1099/mic.0.060855-0
  • Besingi RN, Wenderska IB, Senadheera DB, et al. Functional amyloids in Streptococcus mutans, their use as targets of biofilm inhibition and initial characterization of SMU_63c. Microbiol-Sgm. 2017;163(4):488–501. doi: 10.1099/mic.0.000443
  • Gilbert P, Allison DG, McBain AJ. Biofilms in vitro and in vivo: do singular mechanisms imply cross-resistance? J Appl Microbiol. 2002;92(Suppl (2002)):98s–110s. doi: 10.1046/j.1365-2672.92.5s1.5.x
  • Liu J, Zhang JY, Guo LH, et al. Inactivation of a putative efflux pump (LmrB) in Streptococcus mutans results in altered biofilm structure and increased exopolysaccharide synthesis: implications for biofilm resistance. Biofouling. 2017;33(6):481–493. doi: 10.1080/08927014.2017.1323206
  • Liu J, Guo LH, Liu JW, et al. Identification of an efflux transporter LmrB regulating stress response and extracellular polysaccharide synthesis in Streptococcus mutans. Front Microbiol. 2017;8:962. doi: 10.3389/fmicb.2017.00962
  • Rainey K, Michalek SM, Wen ZT, et al. Glycosyltransferase-mediated biofilm matrix dynamics and virulence of Streptococcus mutans. Appl Environ Microbiol. 2019;85(5). doi: 10.1128/AEM.02247-18
  • Shankar M, Hossain MS, Biswas I. Pleiotropic regulation of virulence genes in Streptococcus mutans by the conserved small protein SprV. J Bacteriol. 2017;199(8). doi: 10.1128/JB.00847-16
  • Inagaki S, Matsumoto-Nakano M, Fujita K, et al. Effects of recombinase a deficiency on biofilm formation by Streptococcus mutans, oral microbiol. Immunol. 2009;24(2):104–108. doi: 10.1111/j.1399-302X.2008.00480.x
  • Zhang J, Liu J, Ling J, et al. Inactivation of glutamate racemase (MurI) eliminates virulence in Streptococcus mutans. Microbiol Res. 2016;186-187:1–8. doi: 10.1016/j.micres.2016.02.003
  • Aleti G, Baker JL, Tang X, et al. Identification of the bacterial biosynthetic gene clusters of the oral microbiome illuminates the unexplored social language of bacteria during health and disease. MBio. 2019;10(2). doi: 10.1128/mBio.00321-19
  • Bushin LB, Clark KA, Pelczer I, et al. Charting an unexplored streptococcal biosynthetic landscape reveals a unique peptide cyclization motif. J Am Chem Soc. 2018;140(50):17674–17684. doi: 10.1021/jacs.8b10266
  • Li ZR, Sun J, Du Y, et al. Mutanofactin promotes adhesion and biofilm formation of cariogenic Streptococcus mutans. Nat Chem Biol. 2021;17(5):576–584. doi: 10.1038/s41589-021-00745-2
  • Guo L, McLean JS, Yang Y, et al. Precision-guided antimicrobial peptide as a targeted modulator of human microbial ecology. Proc Natl Acad Sci U S A. 2015;112(24):7569–7574. doi: 10.1073/pnas.1506207112
  • Lin YW, Gong T, Ma QZ, et al. Nicotinamide could reduce growth and cariogenic virulence of Streptococcus mutans. J Oral Microbiol. 2022;14(1). doi: 10.1080/20002297.2022.2056291
  • Bowen WH, Burne RA, Wu H, et al. Oral biofilms: pathogens, matrix, and polymicrobial interactions in microenvironments. Trends Microbiol. 2018;26(3):229–242. doi: 10.1016/j.tim.2017.09.008
  • Koo H, Sheng J, Nguyen PT, et al. Co-operative inhibition by fluoride and zinc of glucosyl transferase production and polysaccharide synthesis by mutans streptococci in suspension cultures and biofilms. FEMS Microbiol Lett. 2006;254(1):134–140. doi: 10.1111/j.1574-6968.2005.00018.x
  • Binder TP, Robyt JF. Inhibition of Streptococcus mutans 6715 glucosyltransferases by sucrose analogs modified at positions 6 and 6’. Carbohydr Res. 1985;140(1):9–20. doi: 10.1016/0008-6215(85)85045-X
  • Tanriseven A, Robyt JF. Synthesis of 4,6-dideoxysucrose, and inhibition studies of leuconostoc and Streptococcus D-glucansucrases with deoxy and chloro derivatives of sucrose modified at carbon atoms 3, 4, and 6. Carbohydr Res. 1989;186(1):87–94. doi: 10.1016/0008-6215(89)84007-8
  • Young DA, Bowen WH. The influence of sucralose on bacterial metabolism. J Dent Res. 1990;69(8):1480–1484. doi: 10.1177/00220345900690080601
  • Wright WG, Thelwell C, Svensson B, et al. Inhibition of catalytic and glucan-binding activities of a streptococcal GTF forming insoluble glucans. Caries Res. 2002;36(5):353–359. doi: 10.1159/000065962
  • Battagim J, Souza VT, Miyasaka NR, et al. Comparative study of the effect of green and roasted water extracts of mate (Ilex paraguariensis) on glucosyltransferase activity of Streptococcus mutans. J Enzyme Inhib Med Chem. 2012;27(2):232–240. doi: 10.3109/14756366.2011.585986
  • Branco-de-Almeida LS, Murata RM, Franco EM, et al. Effects of 7-epiclusianone on Streptococcus mutans and caries development in rats. Planta Med. 2011;77(1):40–45. doi: 10.1055/s-0030-1250121
  • Koo H, Nino de Guzman P, Schobel BD, et al. Influence of cranberry juice on glucan-mediated processes involved in Streptococcus mutans biofilm development. Caries Res. 2006;40(1):20–27. doi: 10.1159/000088901
  • Nakahara K, Kawabata S, Ono H, et al. Inhibitory effect of oolong tea polyphenols on glycosyltransferases of mutans Streptococci. Appl Environ Microbiol. 1993;59(4):968–973. doi: 10.1128/aem.59.4.968-973.1993
  • Eguchi Y, Kubo N, Matsunaga H, et al. 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(4):1475–1484. doi: 10.1128/AAC.01646-10
  • Muras A, Mayer C, Romero M, et al. Inhibition of Steptococcus mutans biofilm formation by extracts of tenacibaculum sp. 20J, a bacterium with wide-spectrum quorum quenching activity. J Oral Microbiol. 2018;10(1):1429788. doi: 10.1080/20002297.2018.1429788
  • Opoku-Temeng C, Zhou J, Zheng Y, et al. Cyclic dinucleotide (c-di-GMP, c-di-AMP, and cGAMP) signalings have come of age to be inhibited by small molecules. Chem Commun. 2016;52(60):9327–9342. doi: 10.1039/C6CC03439J
  • Rao F, See RY, Zhang DW, et al. YybT is a signaling protein that contains a cyclic dinucleotide phosphodiesterase domain and a GGDEF domain with ATPase activity. J Biol Chem. 2010;285(1):473–482. doi: 10.1074/jbc.M109.040238
  • Yan JC, Gong T, Ma QZ, et al. vicR overexpression in Streptococcus mutans causes aggregation and affects interspecies competition. Mol Oral Microbiol. 2023;38(3):224–36. doi: 10.1111/omi.12407