We use cookies to improve your experience. By continuing to browse this site, you accept our cookie policy.×

Paradigm shift in discovering next-generation anti-infective agents: targeting quorum sensing, c-di-GMP signaling and biofilm formation in bacteria with small molecules

    ,
    Jacqueline AI Smith

    Department of Chemistry and Biochemistry, University of Maryland, College Park, MD 20742, USA

    ,
    Jingxin Wang

    Department of Chemistry and Biochemistry, University of Maryland, College Park, MD 20742, USA

    ,
    Shizuka Nakayama

    Department of Chemistry and Biochemistry, University of Maryland, College Park, MD 20742, USA

    &
    Lei Yan

    Department of Chemistry and Biochemistry, University of Maryland, College Park, MD 20742, USA

    Published Online:https://doi.org/10.4155/fmc.10.185

    Abstract

    Small molecules that can attenuate bacterial toxin production or biofilm formation have the potential to solve the bacteria resistance problem. Although several molecules, which inhibit bacterial cell-to-cell communication (quorum sensing), biofilm formation and toxin production, have been discovered, there is a paucity of US FDA-approved drugs that target these processes. Here, we review the current understanding of quorum sensing in important pathogens such as Pseudomonas aeruginosa, Escherichia coli and Staphylococcus aureus and provide examples of experimental molecules that can inhibit both known and unknown targets in bacterial virulence factor production and biofilm formation. Structural data for protein targets that are involved in both quorum sensing and cyclic diguanylic acid signaling are needed to aid the development of molecules with drug-like properties in order to target bacterial virulence factors production and biofilm formation.

    Figure 1.  Chemical structures of antibiotics.
    Figure 2.  Chemical structures of recently approved antibiotics.
    Figure 3.  The quorum-sensing processes of various Gram-negative and Gram-positive bacteria.

    (A) Multicomponent QS circuits: (i) multiple different autoinducers, AI-1, CAI-I and AI-2, are synthesized by their respective synthases Lux-I and CqsA types and LuxS; (ii) the autoinducers bind specific membrane-bound receptors, LuxN-Type and LuxP; (iii) the signal is relayed to a response regulator via phosphotransfer reactions; (iv) the response regulator activates or represses the expression of QS related genes. Organisms that use multicomponent QS circuits include Vibrio harveyi[66] and Vibrio cholerae[70,71]. (B) Single component QS circuit: (i) an autoinducer is synthesized by its synthase (LuxI-Type); (ii) it diffuses into the cell and binds its receptor (LuxR-type); (iii) the receptor activates or represses the expression of QS-related genes [66]. (C) Autoinducer processing quorum-sensing circuit: (i) the autoinducer AI-2 is synthesized by its respective synthase, LuxS; (ii) AI-2 is transported into the cell via the transport protein, LsrB. (iii) AI-2 is phosphorylated by LsrK-type kinase; (iv) phospho-AI-2 binds to a transcriptional repressor, LsrR-type; (v) binding of phospho-AI-2 to LsrR-type destabilizes the repressor-DNA complex, leading to the expression of QS-regulated genes. Organisms that use the autoinducer-processing QS circuit include Salmonella typhimurium and Escherischia coli[67]. (D) Inter-kingdom QS circuit: (i) the autoinducer, AI-3, is produced by unknown enzymes. Also, hormones are produced by host cell; (ii) autoinducers and hormones from the host organism are detected by a receptor protein (QseC-Type); (iii) binding to the receptor initiates a signal transduction pathway, which activates a response regulator; (iv) the response regulator activates or represses QS regulated genes. Entereohemorrhagic E. coli uses this type of system [68,69,72]. (E) Modified peptide QS circuit: (i) precursor peptides are cleaved, modified and exported out of the cell; (ii) Modified cyclic peptides bind to receptors on the cell surface; (iii) binding to a receptor initiates a signal transduction that leads to activation of a response regulator; (iv) The response regulator activates a gene operon; (v) Quorum sensing-controlled genes are repressed or activated. Organisms that use the modified peptide QS circuit include S. aureus and E. faecium. (F) i–v are similar to (E), except that AIPs are linear peptides.

    AgrD: precursor peptide; AgrB: Modifying/exporting protein; AgrC and A: AIP receptor and response regulator; QS: Quorum sensing.

    Figure 4.  Naturally occurring autoinducers.
    Figure 5.  Mouse infected with (A) Staphylococcus aureus agrC-1 bacterial strain; (B) S. aureus agrC-1+ strain and (C) S. aureus agrC-1+ strain and treated with synthetic autoinducing peptide-II (25).

    Reproduced with permission from [54].

    © National Academy of Science.

    Figure 6.  Processing of DPD in enteric bacteria.
    Figure 7.  Biosynthesis of (A) acylhomoserine lactones and (B) AI-2.

    DPD (47) is in equilibrium with other cyclic structures, shown in Figure 4.

    Figure 8.  AI-2/AI-1 synthesis inhibitors.
    Figure 9.  Quorum-sensing (QS) inhibitors, which bind QS receptors.

    These examples are not exhaustive and only represent the diverse structures of QS inhibitors. For further examples of QS antagonists, see elsewhere [127,128].

    Figure 10.  3D micrograph of biofilm grown on glasses for 48 h.

    (A) Untreated; (B) treated with indole acylhomoserine lactone (64); and (C) treated with p-bromophenyl-acylhomoserine lactone (63).

    Scale bar = 50 µm.

    Reproduced with permission from [132]. © American Chemical Society (2005).

    Figure 11.  Fluorescence imaging of AI-3-mediated A/E lesion formation in enterohemorrhagic Escherichia coli.

    (A) No treatment; (B) treatment with 5 µM LED209 (72) and (C) treatment with 5 pM LED209 (72).

    EHEC: Enterohemorrhagic Escherichia coli.

    Reproduced with permission from [151]. © American Association for the Advancement of Science (2008).

    Figure 12.  Other inhibitors of quorum sensing.
    Figure 13.  Pseudomonas aeurginosa biofilm in a flow cell attenuated by brominated furanones.

    (A) Incubated with exogeneous 3-oxo-C12-HSL (35) (40 nM); (B) incubated with exogeneous 3-oxo-C12-HSL (40 nM) and 2 µg/ml of compound 79 and (C) incubated with exogeneous 3-oxo-C12-HSL (80 nM) and 2 µg/ml of 79.

    Reproduced with permission from [153]. © Society of Microbiology (2002).

    Figure 14.  Lung tissue of mice infected with Pseudomonas aeruginosa.

    (A) Photomicrographs of lung section from a placebo control mouse indicating a clear lung abscess (shown in circle, ×25) and the polymorphonuclear leucocytes or pus cells along the abscess wall (×400, hematoxylin and eosin staining). (B) Lung section from a furanone (80)-treated mouse indicating that the polymorphonuclear leucocyte infiltration in the lung focus is not significant and the lung tissues exhibit a chronic inflammation (center: ×100, right: ×400, hematoxylin and eosin staining).

    Reproduced with permission from [154]. Oxford University Press (2004).

    Figure 15.  Small-molecule inhibitors of virulence factors (both direct and indirect inhibitors).
    Figure 16.  Small molecule biofilm inhibitors.
    Figure 17.  New antimicrobial agents.
    Figure 18.  Structure of cyclic diguanosine monophosphate.
    Figure 19.  Oxidative damage caused by bactericidal antibiotics.

    Adapted with permission from [236].

    The discovery and development of several antibiotics between the late 1930s and 1960s allowed the effective management of previously lethal bacterial infections [1,2]. Penicillin G (2) dubbed the ‘magic drug,’ was responsible for reducing the number of deaths among soldiers in World War II infected with Streptococcus pneumonia[1,2]. When early resistance to penicillin was observed, second-generation antibiotics, such as methicillin (3), cephalothin (4) and imipenem (5), were developed [2]. In 1961, a methicillin-resistance strain of S. aureus (MRSA) was observed [3]. Soon thereafter other bacterial strains that were resistant to antibiotics such as streptomycin (8), chloramphenicol (9) and tetracycline (6) were identified [1,4–6]. Over the years, it has become clear that bacteria can develop resistance to almost any antibiotic.

    Table 1 illustrates the severity of the antibiotic resistance problem. With the exception of a few antibiotics, such as vancomycin (14) and erythromycin (7), resistance to most antibiotics was observed only a few years after their introduction into clinical use. In the case of vancomycin and erythromycin, it took almost three decades between the year that the drug was introduced in the US market and the year resistance was first observed (Table 1)[1,7–20].

    The three main strategies that are used by bacteria to develop resistance to antibiotics are:

    • ▪ Overexpression of enzymes that can modify the antibiotic drug, rendering the antibiotic inactive;

    • ▪ Mutation of the bacterial target site that allows the target site to maintain its functional role yet abrogates binding of drug to the target or transverse of the antibiotic across the bacterial cell wall;

    • ▪ Export of antibiotic drugs to the extracellular media via multidrug-resistant (MDR) efflux pumps or loss of porin channels resulting in lower permeability of antibiotics.

    Drugs such as penicillin G (2) and methicillin (3; β-lactams, see Figure 1) are inactivated via the overexpression of β-lactamases, which hydrolyze the antibiotics [21]. For an excellent review on the mechanism of action of β-lactamases, see elsewhere [22]. Resistance to β-lactams via β-lactamase-mediated hydrolysis can be ameliorated when the β-lactam antibiotic is used in combination with β-lactamase inhibitors such as clavulanic acid [23], tazobactam [23] and sulbactam [24]. Both clavulanic acid and sulbactam irreversibly inactivate β-lactamases [22–24]. Unfortunately some bacterial strains that are resistant to β-lactamase inhibitors have emerged [25]. Aminoglycosides, including streptomycin (8) and tobramycin, which inhibit bacterial rRNA are rendered ineffective by bacteria via the over-expression of antibiotic-modifying enzymes, such as nucleotidyltransferase/phosphotransferase, which transfer a monophosphate moiety from ATP to a hydroxyl on the ring of an aminoglycoside [26]. Other aminoglycoside modifying enzymes include N-acetyltransferases and O-adenyltransferases, which catalyze acetyl CoA-dependent acetylation of amino functionality and ATP-dependent adenylation of hydroxyl group, respectively [26]. Other antibiotics such as the streptogramin class and linezolid (10) are rendered ineffective via the modification of 23S ribosomal RNA [27]. Bacteria can use multiple pathways to inactivate an antibiotic. For example, resistance to vancomycin (14) can occur via thickening of the bacterial cell wall [28], changing the peptidoglycan precursor [1] or expression of an ‘abnormal’ ligase that makes the unusual Ala-lactate peptide bond [29].

    Efflux pump proteins such as AcrB in E. coli[30], the Mex-Opr systems in P. aeruginosa[31,32] and NorA in S. aureus[33] export antibiotics such as ciprofloxacin (13)[33] and tetracycline (6)[34,35] out of bacterial cells, thereby raising the minimum inhibitory concentration values of antibiotics. Inhibitors of MDR efflux pumps such as the alkaloid reserpine, biricodar and timcodar (which are also mammalian multidrug efflux inhibitors) have been shown to potentiate antibiotic activity of conventional antibiotics [36]. Other small molecules such as L-phenylalanine-L-arginine-β-naphthylamine and carbonyl cyanide m-chlorophenylhydrazone (CCCP) have also been shown to inhibit the efflux of drugs such as levofloxacin and faropenam from bacterial cells [37,38]. The widespread use of MDR efflux inhibitors as combination therapeutics is probably not as popular as the use of β-lactamase inhibitors, due to the paucity of effective bacterial MDR efflux pump inhibitors.

    After years of wavering success with the development of second- and third-generation antibiotics that would ultimately suffer from bacterial resistance, two new drugs were introduced at the beginning of the new millennium; daptomycin (15)[39,40] and linezolid (10)[41,42]. Linezolid (10) is a synthetic oxazolidinone antibiotic, which was initially effective against vancomycin-resistant Enterococci (VRE) and MRSA [41,42]. Linezolid inhibits protein synthesis by binding to the 23S site of the 50S ribosome [41,42]. Although its action is similar to the macrolides and phenylpropanoids, no cross-resistance was observed [41,42]. Regretfully, linezolid resistance was reported in VRE and MRSA only a few years after it was introduced [19,43]. Daptomycin (15), a natural lipopeptide antibiotic, was also introduced in the early 21st century [39,40]. Daptomycin kills bacteria via a novel mode of action involving binding to the bacterial cell wall and depolarizing the cell membrane [39,40]. Unfortunately, the early excitement regarding daptomycin’s ability to kill MRSA and VRE was diminished upon the identification of strains of MRSA and VRE that were resistant to daptomycin (15)[44].

    Only a handful of antibiotics have been approved by the US FDA for clinical use in the past few years (see Table 2 and Figure 2) [45–47]. Worryingly, these newer antibiotics (with the exception of retapamulin) are merely derivatives of older generation antibiotics. Since the newer antibiotics are acting on the same targets, which have been shown to be evolvable, it would be naive to think that the newly introduced antibiotics will not ultimately suffer from bacterial resistance.

    Retapamulin (21), introduced in 2007, belongs to a new class of antibiotics called pleuromutilins. Retapamulin inhibits protein biosynthesis by binding to the peptidyl transferase center of the 23S rRNA, which is located on the 50S subunit and, thus, prevents peptide bond formation [48]. This mode of action differs from the macrolide protein inhibitors such as erythromycin (7) and azithromycin, which also bind to the 23S rRNA segment of the 50S subunit, but do not interrupt the peptidyl transferase [48]. Retapamulin (21) may also destabilize tRNA in the P-site of the 50S subunit [48]. However, bacteria have shown in the past that ribosomal targets can be modified in order to abrogate antibiotic inhibition and there is no reason to suggest that the 50S ribosome is not evolvable.

    The previous discussion highlights an unsavory truth: we are in a never-ending battle against pathogenic bacteria. The ability of bacteria to find several pathways to render a bacteriocidal or bacteriostatic antibiotic ineffective, coupled with the lack of newer druggable targets in bacteria and the paucity of new structural classes of antibiotics, implies that there might be a need for a paradigm shift in the strategies used to treat bacterial infections.

    Over the last few years, researchers have begun to focus attention on finding ways to render pathogenic bacteria less virulent with small molecules. It is assumed that if one can find ways to prevent bacteria from being virulent (i.e., inhibit toxin production) without necessarily killing the bacteria, then there will be less evolutionary pressure for the bacteria to evolve resistance mechanisms. However, this presupposes that toxin production by bacteria does not confer a competitive advantage. Since biofilms are difficult to treat with conventional antibiotics, it is also expected that small molecules that can either inhibit biofilm formation or disperse an existing biofilm will find utility in combination antibiotic therapy. Studies from several laboratories have unraveled the complexities of the signal transduction pathways that lead to both toxin production and biofilm formation in bacteria. It is now known that the majority of bacteria toxin production is quorum sensing (QS)-controlled, vide infra[49]. Futhermore, a dinucleotide known as cyclic diguanylic acid (c-di-GMP) appears to be a master regulator of diverse bacteria strains [50,51]. In this article, the various signaling transduction processes in bacteria that result in virulence production or biofilm formation are briefly discussed and the small molecules that have been used to attenuate bacterial virulence as well as biofilm formation will be examined. Space limitation prevents us from discussing all of the molecules that have been shown by several groups to make bacteria avirulent. Therefore, the molecules that are chosen in this review are merely representative examples and are in no way exhaustive.

    Quorum sensing

    Quorum sensing is a form of bacterial cell-to-cell communication whereby bacteria secrete and detect signaling molecules known as autoinducers. The detection of a threshold concentration of autoinducers induces the activation/suppression of genes responsible for processes such as virulence gene expression [52–60], biofilm formation [61–64] and antibiotic resistance [65]. Different QS circuits have been identified. Some bacteria use multiple QS circuits to monitor their environment. Figure 3 illustrates six general QS systems described to date [66–72].

    Bacteria use several different classes of molecules for cell-to-cell communication (Figure 4). Several Gram-positive bacteria use peptides (25–28) for intraspecies communication [73] whereas many Gram-negative bacteria utilize acylhomoserine lactones (AHL; also called AI-1; 29–39) [56,66,74] for intraspecies communication. Other molecules, such as g-butyrolactones (43), quinolones (40) and hydroxy ketones (42), are also employed as intraspecies communication molecules in Streptomyces[75], Pseudomonas aeruginosa[76] and Vibrio cholerae[70,71], respectively. AI-2 (collective name for the family of interconverting cyclic lactols 44–46; Figure 4) has been described as being used by bacteria for interspecies communication [67]. AI-2 has been found in several species of bacteria [77,78] and is derived from 4,5-dihydroxyl-2,3-pentadione (DPD; 47) [79,80]. Although over 70 bacterial strains have been shown to harbor the enzyme that makes AI-2, known as LuxS, only a handful of bacterial species have been shown to possess receptors that sense and respond to the presence of AI-2; the LuxP receptor in Vibrio harveyi and LsrB in E. coli/Salmonella bind to different forms of AI-2; LuxP binds to furanosyl borate diester [79] whereas LsrB binds to (2R,4S)-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (R-THMF), which lacks boron [80]. Another protein that binds ‘processed’ AI-2 is the LsrR transcriptional regulator in E. coli/Salmonella that binds to phosphorylated AI-2 [67,81]. Despite the paucity of identified bona fide AI-2 receptors in bacteria, the role of this molecule as a cell-to-cell signaling molecule and mediator of virulence production in some bacteria is now established. For example, in Vibrio vulnificus[82,83], Serratia marcescens[84] and Clostridium perfringens[85], AI-2 has been implicated in the production of virulence factors. However, the receptors that mediate virulence-factor production in the aforementioned bacteria have yet to be fully characterized. In other bacteria, LuxS (the synthase for AI-2) might have an additional role in the metabolic pathway involved in methionine metabolism [77,86]. The LuxS-independent formation of AI-2 from ribulose-5-phosphate has been proposed [87–89]. The significance and implications of alternative AI-2 synthesis in QS remains to be fully elucidated.

    Bacteria can also sense molecules produced by eukaryotic host cells and regulate genes in response to molecules produced by the host. For example, enterohemorrhagic E. coli (EHEC) serotype O157:H7, which is responsible for outbreaks of bloody diarrhea and uremic syndrome throughout the world, responds to the eukaryotic hormone epinephrine. β- and α-adrenergic antagonists have been shown to inhibit the bacterial response to this hormone [68]. EHEC and several other bacteria also produce an autoinducer known as AI-3 [69], whose structure still remains to be elucidated.

    ▪ QS in a select few human pathogens

    Quorum sensing systems in Pseudomonas aeruginosa, Staphylococcus aureus, Vibrio cholerae, Salmonella typhimurium and Escherichia coli have been of interest to several researchers due to numerous common diseases that are attributed to these opportunistic bacteria. The QS circuitry in P. aeruginosa, a Gram-negative bacteria that thrives in many environments, has been well studied. P. aeruginosa can cause infections of the respiratory, urinary and GI tracts, as well as keratitis and bacteremia. Cystic fibrosis, burn or immunocompromised patients are particularly susceptible to P. aeruginosa infections. Two major QS systems in P. aeruginosa are the las and rhl systems [90]. The las and rhl QS systems are composed of transcriptional regulators that respond to autoinducers produced by LasI and RhlI synthases. LasI produces 3-oxo-C12-HSL (35; Figure 4), which is sensed by LasR, whereas RhlI produces C4-HSL (29), which is sensed by RhlR. The las and rhl QS systems are organized in a hierarchical manner; LasR: 3-oxo-C12-HSL complex regulates the expression of the rhl system. Also, it is believed that the las system regulates the rhl system post-translationally. This is achieved via the competitive binding of 3-oxo-C12 HSL to RhlR at low C4-HSL concentration [90]. A transcriptional regulator, QscR represses the transcription of lasI and rhlI as well as lasB, pqsH, hcnAB and phenazine genes, phzA and phzA2[91,92]. P. aeruginosa also possesses a third QS molecule known as the Pseudomonas quinolone signal (2-heptyl-3-hydroxy-4-quinolone (40)) [76]. Microarray analysis has shown that a few hundreds of genes are under QS control in P. aeruginosa[93–95]. These genes regulate diverse toxic factors such as elastase (lasB) [96], LasA protease (lasA) [97], alkaline proteases (apr) [98] and secretion pathway (xcpP and xcpR) [99]. A fourth class of molecules found in P. aeruginosa is the dipeptide-derived piperazines (41)[100]. P. aeruginosa may use these piperazines to communicate with other bacteria such as P. mirabilis, C. freundii and E. agglomerans. More work is needed to confirm whether dipeptide-derived piperazines are indeed QS signaling molecules [101].

    S. aureus is an opportunistic pathogen and is the major cause of postoperational infections as well as pneumonia, acute endocarditis and abscesses of the skin [52–54]. The accessory gene regulator (agr) QS system in S. aureus and other staphylococcal species plays a central role in the regulation of virulence factors during staphylococcal pathogenesis. The agr locus consists of the P2 operon (agrACDB) and the P3 operon, whose transcript, RNAIII encodes α-toxin (hld), in addition to having regulatory effects on agr expression [102]. In the agr QS system, the autoinducing peptide (AIP), consisting of seven to nine residues, is processed from AgrD protein via AgrB-mediated thiolactone ring formation between a conserved central cysteine and a C-terminal carboxyl group [53]. It is plausible that other proteins apart from the membrane bound AgrB might be involved in AgrD processing. The secreted AIP is recognized by the membrane-bound AgrC and triggers the transfer of a phosphate group between AgrC and AgrA. AgrA together with other proteins such as the transcriptional regulator SarA activates the transcription of the P2 and P3 operons [103]. The agr system regulates an array of virulence factors. These include entereotoxin B and C, proteases (splA, B, D, F and V8) and cell-surface associated capsular polysaccharides [104,105]. Interestingly, the AIP produced by one of the four distinct agr groups generally inhibits signaling by another Staphylococci that harbors a different agr group. The inhibition occurs via competitive binding to the AgrC receptor. For example, an AIP produced by S. epidermidis can inhibit all but type IV of S. aureus agr groups [52]. Although recent evidence suggest that agr expression in vivo is more complex than originally thought and might depend on several environmental factors [106], it is clear that inhibition of the agr QS system will attenuate S. aureus disease progression and persistence. This is illustrated in Figure 5. In Figure 5A, a mouse was infected with S. aureus that lacked the AgrC-1 receptor, whereas in Figure 5B, the mouse was infected with S. aureus that was agrC-1+. The mouse that was infected with S. aureus (agrC-1+) developed an abscess that was significantly larger than the mouse that was infected with S. aureus (agrC-1). Furthermore, when the mouse was infected with S. aureus (agrC-1+) and then treated with AIP-II (25b), the abscess that developed was similar to the agrC-1 infection case. This showcases that the inhibition of one group of agr by AIP produced by another group also occurred under in vivo conditions. Other Gram-positive bacteria also use peptides for QS: E. faecium uses GBAP (28)[73], S. pneumoniae uses competence stimulating peptide (CSP; 26) [107] and B. subtilis uses competence and sporulation factor (CSF) and ComX (27)[66,107,108].

    Another human pathogen, V. cholerae causes the deadly disease cholera [70]. V. cholerae is a Gram-negative bacterium and uses two signaling molecules for QS: (S)-3-hydroxytridecan-4-one (or C-AI-1 42) and AI-2 (44–46) [71]. (S)-3-hydroxytridecan-4-one (42) is the major QS molecule used in V. cholerae and it is structurally unique [71]. Compound 42 is a 13-carbon hydroxyl ketone (Figure 4) and is synthesized by CqsA and detected by CqsS [71,109]. Ironically, the QS circuitry is reversed in V. cholerae; expression of virulence and biofilm formation is activated at low cell density and repressed at high cell density [70,71]. The repression of biofilm formation and virulence expression at high cell density allows the bacteria to leave the host and be re-introduced into the environment [70,71].

    Salmonella typhimurium is the leading cause of food-borne illnesses worldwide and causes gastroenteritis. Although S. typhimurium poisoning (‘food poisoning’) often clears on its own, it can be fatal in immunocompromised patients. In addition, biofilm formation on abiotic surfaces (e.g., restaurant kitchen counters) is a major cause of the spread of diseases. QS in S. typhimurium utilizes AHLs and AI-2 and their receptor proteins SdiA and LsrB, respectively [58–60]. SdiA detects acylhomoserine lactones, although S. typhimurium does not produce AHLs on its own [58–60]. The detection of AHLs is thought to be a defense mechanism that S. typhimurium uses to survive in environments with other diverse pathogens [58–60]. In S. typhimurium and E. coli, when a threshold concentration of AI-2 is reached, LsrB transports the molecule inside the cell where it is phosphorylated by LsrK kinase [67]. Phospho-AI-2 (48; Figure 6) binds and de-represses the LsrR repressor protein, which controls the lsr operon [67,81]. Ultimately, LsrG protein degrades phospho-AI-2 (48) into phosphoglycolic acid (PG; 49) and an unknown C3 compound (Figure 6)[67]. It remains unclear why Salmonella and E. coli import and subsequently degrade AI-2, although it is plausible that AI-2 transport and degradation is a defense mechanism to quench the AI-2 signal of other bacteria [110].

    ▪ Interception of QS signaling:

    If QS is vital for the production of bacterial virulence, then it is reasonable to suggest that any of the following strategies will attenuate bacterial virulence production:

    • ▪ Inhibition of enzymes that catalyze autoinducer synthesis;

    • ▪ Inhibition of receptors (both cell surface and cytosolic) that bind to autoinducers or processed autoinducers;

    • ▪ Inhibition of downstream proteins that link QS receptor occupancy to activation of QS-regulated genes via phosphorelay mechanisms;

    • ▪ Inhibition of virulence factors that are expressed in response to QS.

    In this article, we will discuss the nature of the small molecules that have been shown to inhibit QS at the various stages. Owing to space limitations, we are unable to list all of the chemical entities that have been shown to antagonize QS in pathogenic bacteria. We have only selected examples that have reasonable potency in attenuating QS processes. The majority of small molecules with EC50/IC50 values greater than or equal to 50 µM are excluded from this article.

    ▪ Inhibition of autoinducer biosynthesis

    Acyl-homoserine lactones used for signaling in Gram-negative bacteria are synthesized by the reaction of the acyl group located on an acyl-carrier protein with S-adenosyl-methionine (50)[111]. Briefly, the acyl group on the acyl-carrier protein is transferred to the amino functionality in SAM (50). An intramolecular nucleophilic substitution reaction involving the carboxylate moiety of SAM and the thioether as a leaving group leads to the formation of AHLs and methylthioadenosine (51)[111]. The nucleosidase, Pfs, transforms methylthioadenosine (51) into methylthioribose (52) by cleaving the adenosine group. Methylthioribose is further degraded into methionine (53), which is recycled into the SAM biosynthesis pathway (Figure 7A). AI-2 biosynthesis also uses SAM as its starting point. In AI-2 biosynthesis, a methyltransferase removes the methyl group from SAM; converting it to S-adenosylhomocysteine (SAH; 54) [112]. Similar to the AHL synthesis, Pfs then cleaves the adenosine group of SAH (54) to form S-ribosylhomocysteine (SRH;55) [112]. Finally, LuxS converts SRH into DPD (47) and homocysteine (56; Figure 7B) [112].

    Since AI-2 synthase, LuxS, has been found in over 70 bacterial species [77,78], inhibitors of LuxS could have broad-spectrum anti-QS effects. On the other hand, AI-1 synthases are species specific, therefore it might be difficult to find broad-spectrum AI-1 synthase antagonists. Attempts to utilize SAM analogs to inhibit AI-2 synthesis have been described [113], but it is unclear if this is a viable strategy since SAM is vital for several processes in both bacterial and human cells. On the bright side, inhibition of methylthioadenosyl nucleosidase, Pfs, involved in AI-1 and AI-2 biosynthesis has shown great promise. Transition state analog inhibitors of 5´-methylthioadenosine/S-adenosylhomocysteine nucleosidase (58; Figure 8) were found to inhibit 5´-methylthioadenosine/S-adenosylhomocysteine nucleosidase with Ki values as low as 47 fM [114]. More excitingly, these analogs (58a) have been shown to be effective against biofilm formation in a variety of bacteria such as V. cholerae and E. coli O157:H7 [115]. LuxS inhibitors are also promising drugs and several have been successfully synthesized [116–120]. One of the most effective LuxS inhibitors reported to date is compound 57(Figure 8). This compound was found to inhibit LuxS of B. subtilis with a Ki value of 0.37 µM [118]. Compound 57 had less affinity for the LuxS of E. coli and V. harveyi (Ki∼13 µM for both enzymes) [118]. Although all LuxS enzymes catalyze the formation of AI-2, subtle differences amongst LuxS from different bacterial strains might preclude the use of a single chemical entity to inhibit all LuxS enzymes.

    ▪ Inhibition of QS by targeting receptor proteins

    Several research groups have pursued the development of novel QS inhibitors that target receptor proteins that bind autoinducers [121–132]. Since QS receptors might not be found in humans, targeting these proteins could be less cytotoxic than other anti-QS strategies. It is, however, plausible that eukaryotic hosts have receptors that sense bacterial signaling molecules and the receptors are yet to be found [133]. The autoinducer-binding domain of the LuxR-family of proteins resembles that of the GAF or PAS domain that is present in a number of mammalian signaling proteins and transcriptional factors [133]. AI-1 from P. aeruginosa has been shown to affect the production of cytokines by immune cells and also increase IL-8 mRNA levels in epithelial cells [134]. Several anti-QS small molecules have been developed that are structurally similar to natural autoinducers (Figure 4). Among Gram-positive bacteria, a cyclic peptide analog of AIP-I (25a), AIP-I D5A (60; Figure 9), has been shown to effectively inhibit AgrC3, while AIP-I D5N and AIP-IV Y5F were shown to inhibit AgrC4 and AgrC2, respectively, in S. aureus(Table 3)[121,122]. Several AIP analogs that lack a peptide tail have been synthesized and some of these truncated analogs act as inhibitors of AgrC. For example, trAIP-I D2A (62; truncated form of AIP-I) potently antagonizes AgrC1 (Table 3)[121,122].

    The main functional groups that constitute AHLs are a lactone core, an amide moiety and an alkyl side chain. AHLs containing 3-oxo-amide side chains are also found (see compounds 35–39, Figure 4). Anti-QS small molecules based on the AHL structure with one or all of the three core motifs modified have been pursued. For example, the bromophenyl AHL compound (63; the alkyl side chain of this AHL analog contains a para-bromoethylbenzyl group, Figure 9) inhibits LasR, a LuxR-type protein found in P. aeruginosa, at micromolar concentrations [123]. Other AHL analogs, including compounds 64–70, inhibit QS-mediated processes in P. aeruginosa, A. tumefaciens or V. fischeri(Table 3)[123–126].

    LuxR-type proteins bind to AHLs resulting in transcriptional response. AHL autoinducers bind to the amino-terminal module of LuxR-type proteins and, upon autoinducer binding, LuxR-type proteins dimerize and bind to specific DNA sequences near target promoters [135]. Schuster and Greenberg have proposed three different classes of LuxR-type proteins based on biochemical characterization [135]. These classes are:

    • ▪ Class I proteins include LasR and CepR of Pseudomonas aeruginosa[136] and Burkholderia cenocepacia, respectively [137]. These proteins bind to AHLs in an unfolded state but, after AHL binding, class I proteins fold to bind AHL tightly;

    • ▪ Class II proteins include the prototypical LuxR from V. fischeri[138] and QscR from P. aeruginosa[139]. These proteins also require AHL binding for folding but binding to the folded protein is not as tight as it is in the case of class I (i.e., AHL binding to class II is reversible);

    • ▪ Class III proteins include RhlR from P. aeruginosa[140] and ExpR from Erwinia chyrisanthemi[141], which do not require AHL binding for folding. Binding of AHL to class III proteins is reversible; similar to binding to class II proteins.

    Binding of AHLs to LuxR-type receptors results in the expression of virulence factors in several bacteria. For example, during infection by P. aeruginosa and Burkholderia cenocepacia (bacteria that usually coinfect patients with cystic fibrosis), QS-regulated virulence factors such as elastase (lasB), protease A (lasA), exotoxin (toxA) and alkaline protease (aprA) produced by P. aeruginosa[142], and extracellular proteases and siderophores produced by B. cenocepacia lead to fatal pulmonary decline [143]. LuxR-type receptors have also been implicated in bacterial biofilm formation. For example, compounds 63 and 64 attenuate the biofilm formation ability of P. aeruginosa(Figure 10)[132].

    Despite the demonstration that AHL analogs such as 63 can successfully antagonize native AHL-induced virulence and biofilm formation in bacteria, it remains to be shown whether these analogs can survive lactonases that are capable of cleaving the lactone structure under in vivo conditions [144,145]. A recent finding by Greenberg et al. that lactonases produced in the human airway might serve as bacterial QS quenchers [146–148] cautions against the pursuit of lactone or thiolactone-based AI-1 analogs as anti-QS drugs. AHL-based analogs that have the lactone moiety replaced by cyclic structures such as cyclohexanone (69, Figure 9) and phenol (70) have also been pursued. Suga et al. have shown that analogs such as 68, 69 and 70, which are presumably lactonase resistant, can inhibit QS-regulated lasI and rhlI gene expression in P. aeruginosa[125]. Compounds 68, 69 and 70 caused more than 50% reduction in the expression of GFP reporter gene induced by 1 µM AI-1 at micromolar concentrations, providing encouragement that effective lactonase-resistant AHL-based anti-QS drugs could be developed in the future. PD12 (73) contains modifications of the lactone ring as well as the acyl side chain. PD12 (73) is a potent inhibitor of AHL-mediated QS with an IC50 in the nanomolar range (Table 3)[149]. Another AHL analog that contains modification to the acyl side chain is the sulfide 67, which inhibits LasR monitor system with low micromolar IC50[150].

    AI-2 is important in bacterial QS but, despite its widespread utility, very few analogs based on the universal QS molecule AI-2 have been reported as anti-QS agents. Recently, the laboratories of Janda, Meijler and Sintim reported that C1-alkyl analogs of AI-2 did not have agonist activities but rather enhanced AI-2 agonism in V. harveyi’s bioluminescence (synergistic agonism) [129–131]. The C1-butyl analog of AI-2 (71) was shown to inhibit AI-2-induced β-galactosidase activity in S. typhimurium Met-844 with an IC50 value of approximately 5 µM when tested against 50 µM AI-2 [129]. The exact mode of inhibition by the C1-butyl analog of AI-2 is unknown. It is, however, tempting to speculate that this particular analog inhibits LsrR. LsrR is a transcriptional repressor of the lsr operon in enteric bacteria. LsrR binds to phospho-AI-2 and this binding prevents LsrR from being a repressor [67,83].

    QseC is a receptor for AI-3 and is found in EHEC [68]. Recently, high-throughput screening has identified LED209 (72; Figure 9) as a potent inhibitor of QseC [151]. Inhibition of QseC by 72 blocks the expression of virulence factors required for attaching and effacing lesions by EHEC (Figure 11).

    Naturally occurring brominated furanones produced by algae and analogs (79–82, Figure 12) are inhibitors of QS and biofilm formation in a number of bacteria, including P. aeruginosa and S. typhimurium[152–154]. Compounds 79–81 are synthetic analogs of the natural furanone 82. Compound 79 inhibits P. aeruginosa biofilm formation (Figure 13)[153] and compound 80 significantly reduces the formation of abscess on mice lung tissue (Figure 14)[154]. Compound 81 is less potent than the natural furanones (e.g, 82) at reducing biofilm formation in S. typhimurium(Table 3) but is also less toxic [152]. In the S. typhimurium biofilm study, it was shown that flagellar biosynthesis was affected by brominated furanones but none of the known target genes of the AI-2 system and AHL receptor SdiA were affected. It therefore appears that brominated furanones affect biofilm formation in some bacteria by interfering with QS systems, for example in P. aeruginosa[153], whereas in other bacteria, alternative targets might be involved. The anti-QS activity of the halogenated furanones has previously been ascribed to inhibition of the transcriptional regulatory protein, LuxR-type. For example, in V. harveyi, the DNA-binding activity of LuxR is inhibited by the natural furanone (5Z)-4-bromo-5-(bromomethylene)-3-butyl-2(5H)-furanone [155]. A recent study has shown that halogenated furanones might also act to inhibit QS by inactivating AI-2 synthase, LuxS, via a covalent modification [156].

    Several other inhibitors of QS have been discovered using high-throughput screening of both pure compounds and mixtures. Microarray analysis revealed that the addition of 4-nitropyridine-N-oxide (75; Figure 12) downregulates 37% of the QS regulon in P. aeruginosa including virulence factors such as lasB, rhlAb, aprA and phzAIBI[157]. Similarly, patulin (74) and penicillic acid (76) from the broth of Penicillium sp were shown to downregulate 45% and 60% of QS-controlled genes, respectively, in P. aeruginosa[158]. Similarly, GC-7 (77), an extract from garlic, inhibits LuxR but not LasR monitor systems [150]. Another extensive study looked at over 200 marine extracts in order to find potent QS inhibitors [159]. From this study, manoalide derivatives were identified as QS inhibitors. Extracts of the sponge Leucosolenia variabilis were also active QS inhibitors and inhibited lasB expression in P. aeruginosa[159]. The common food additive cinnamaldehyde (78) has been shown to inhibit AI-2 mediated QS by destabilizing the transcriptional regulator LuxR in V. harveyi[160].

    ▪ Macromolecules as anti-QS drugs

    So far we have focused on small molecules that inhibit bacterial QS. The use of protein-based or oligonucleotide-based therapeutics has also been explored as an anti-QS strategy. It has been demonstrated that monoclonal antibodies raised against 3-oxo-C12 AHL or AIP-IV (25d) haptens were able to inhibit AHL- and AIP-mediated QS [161–164]. Monoclonal antibodies for AIP could protect mice from lethal P. aeruginosa infections [165] and inhibit biofilm formation in S. epidermidis[166]. In the natural environment, AHL lactonases (AiiA) and acylases have been shown to degrade AHL and quench AHL-mediated QS [167,168]. Stable versions of these enzymes and other QS-processing proteins could, in principle, be used as protein-based anti-QS therapeutics [169].

    ▪ Virulence factor inhibition

    A high proportion of bacterial virulence factors are under QS control and, hence, inhibition of autoinducer synthases or receptors can reduce virulence factor production, as has been demonstrated in the preceding section of this article. An alternative strategy to curb bacterial virulence is to directly inhibit the factors that are detrimental to the host cell. These factors can be sub-divided as:

    • ▪ Bacterial adhesion molecules that aid the attachment of bacteria to host cell;

    • ▪ Type I–VI secretion systems that provide a channel for the delivery of toxic proteins or peptides from bacterial cells into host cells or the extracellular space;

    • ▪ Toxic factors, which directly disable the normal cellular functions of the host cells.

    Adhesion to host cells is usually a prerequisite for the successful establishment of infection by pathogenic bacteria. Adhesins, retractile type IV pilli or other surface macromolecules are typically used to establish host–pathogen crosstalk that leads to disease progression. Examples of adhesion molecules are the lectin-like E. coli, FimH [72], fibronectin-binding proteins A and B in S. aureus[105] and intimin in EHEC [72]. Since bacterial adhesion is mediated by interactions of carbohydrates that are present on the host cell surface, the development of carbohydrate-based anti-adhesion drugs holds promise [170,171]. Inspiration for the development of carbohydrate-based molecules as anti-adhesion therapeutics comes from the observation that the protective properties of human breast milk are derived from the presence of oligosaccharides [172] . For an excellent review on the anti-adhesive properties of milk oligosaccharides, see elewhere [172,173].

    Compound 85 is a polyvalent inhibitor of anthrax toxin [173]. It blocks the host receptors that are used by B. anthracis to gain entry into the host cell. Compound 85 represents an important antibiotic class that might not suffer from bacterial resistance. This is because extensive alterations to the pathogen’s toxin has to be made in order to switch to another host receptor [173]. Synthetic carbohydrate-based compounds such as 84(Figure 15) have been shown to be effective inhibitors of fimbriae and FimH-mediated adhesion to guinea pig intestinal cells with 700-times higher affinity than methyl-mannoside [174]. Compound 83 is also carbohydrate based and it has been shown to inhibit the colonization of H. pylori in the GI tract of monkeys [175]. Carbohydrate-based compounds found in cranberry juice inhibit fimbriae-facilitated adhesion of E. coli to the urinary wall [176]. Other noncarbohydrate-based compounds such as 86(Figure 15) have been shown to be effective anti-adhesive compounds [177]. Compound 86 inhibits sortase A with a Kiapp of 0.3 µM. Sortase A is used by bacteria, such as S. aureus, for attachment to host cell wall [177]. The pilicide BibC10 (87) inhibits the formation of pili and surface adhesion fibers, in ureopathogenic E. coli[178,179]. BibC10 (87) blocks the usher–chaperone pathway by binding the chaperone, PapD [178,179].

    Bacterial secretion systems are elaborate apparatus or systems that allow bacteria to secrete diverse molecules such as proteins, oligosaccharides and nucleic acids into the extracellular media or directly into host cells. These secretory apparatus thereby play important roles in bacterial pathogenesis. There are six well-known secretion systems in Gram-negative bacteria [180]. These systems secrete proteins either in a single step or with initial export into the periplasm via the Sec pathway and then use other machinery to cross the outer membrane [180]. The Sec system recognizes proteins for transport via a hydrophobic N-terminus signal sequence and then uses ATP to move the protein across the plasma membrane [180]. Type I secretion system (T1SS) is composed of the ABC transporter protein complex, membrane fusion protein and outer membrane protein (OMP). T1SS is important for the secretion of diverse biomolecules ranging in sizes (10–900 kDa) and substrate type (from proteins to polysaccharides). Several important bacterial toxins, such as the pore-forming RTX toxin hemolysin, which kill human immune and other cells are secreted to the exterior via T1SS [180]. The type II secretion system utilizes the Sec system for entry into the periplasm [181]. Once in the periplasm, proteins pass through the outer membrane via a secretion apparatus that is made of a multimeric complex of secretin and other membrane proteins. The type III secretion system (T3SS or TTSS) acts like a needle that penetrates and injects proteins into the host cell [182]. Type IV secretion system (T4SS or TFSS) is capable of transporting DNA and proteins into host cells [183]. Several pathogenic bacteria use the T4SS to deliver toxins such as pertussis toxin (from Bordetella pertussis) and CagA (from Helicobacter pylori). Type V secretion system (T5SS, also known as autotransporter system) also uses the Sec system to cross the membrane. Proteins, which use the type V secretion system, form a β-barrel with the C-terminus and inserts into the outer member [184]. The type V secretion proteins are then cleaved and the β-domain is released [184]. The type VI secretion system proteins (T6SS) presumably does not use the Sec system as they lack N-terminal signal sequences. It is now believed that T6SS are widespread in Gram-negative bacteria [185,186].

    Quorum sensing has been implicated in the formation of bacterial secretion systems [187–189]. Therefore, general inhibitors of the QS circuitry might also inhibit bacterial secretion systems. Imine 89, at micromolar concentrations, has been shown to be an effective inhibitor of T3SS in enteropathogenic E. coli[188]. Compound 89 inhibits T3SS without affecting motility and flagellin expression in enteropathogenic E. coli[188]. Other T3SS inhibitors include compounds 90 and 91, which inhibit T3SS secretion at micromolar concentrations in Yersinia pseudotuberculosis[190]. The exact molecular targets of 89–91 are poorly characterized and, although they are not direct inhibitors of T3SS, they inhibit the expression of virulence-associated factors.

    Compounds that inhibit transcription regulators of toxin genes or that inhibit the toxins directly have also attracted interest. For example, virstatin (93), identified via a high-throughput phenotypic screen, inhibits the transcription regulator ToxT in V. cholerae[191]. By inhibiting ToxT, virstatin (93) prevents the expression of cholera toxin and toxin coregulated pilus. These toxins are responsible for secretory diarrhea and V. cholerae attachment to human intestinal wall [191]. Compound 88 and peptoid 92 have been shown to bind to shiga [192] and cholera toxins [193], respectively. Shiga toxin, produced by Shigella dysenteriae type 1 and Shiga-like toxins from E. coli can cause serious illness to humans. These toxins belong to the bacterial AB5 toxins, which are made up of an enzymatic A subunit that is responsible for the toxin’s action and a homopentamer B5 subunit that is involved in oligosaccharide recognition of the host cells [194].

    RNAIII-activating protein is responsible for a-toxin and serine protease expression in S. aureus[105]. The RNAIII-activating protein inhibitor peptide (RIP) inhibits RNAIII-activating protein and prevents biofilm formation and a-toxin and serine protease expression in S. aureus[195]. However, the potential for enzymatic degradation of RIP by peptidases in vivo could limit the therapeutic potential of RIP [195]. Hamamelitannin (95) also inhibits RNAIII-activating protein [195]. It has, however, not been demonstrated whether the ester functionalities in 95 will survive in vivo esterases. Compound 94, a metabolite extracted from Flustra foliacea, suppresses AHL-dependent production of proteolytic enzymes in P. aeruginosa at micromolar concentrations [196].

    ClpP proteases have been identified as important virulence factors in S. aureus[197]. Clp proteolytic complexes degrade accumulated and misfolded proteins during pathogenesis by S. aureus[198]. Compound 96 has been identified as a micromolar inhibitor of ClpP and effectively prevents haemolysis and proteolysis in S. aureus[197].

    ▪ Biofilm inhibitors

    Over 65% of hospital infections are caused by biofilm-forming bacteria [62–64]. Biofilms are micro-organisms embedded in a matrix and attached to surfaces [62–64]. Once incorporated into the biofilm community, bacteria are resistant to antibiotic action [62–64]. In addition, the biofilm environment increases the probability of antibiotic-resistant plasmid transfer between bacteria [62–64]. Despite the central role that bacterial biofilm plays in infection, there is currently no antibiofilm drug in clinical use. This is probably due to the fact that the mechanisms that underline biofilm formation are not completely understood. There is a paucity of small molecules that can inhibit biofilm or disperse mature biofilm.

    Several of the QS inhibitors, such as brominated furanones (79–82), have been shown to inhibit biofilm formation [152–154]. AHL-analogs (63, 64, 69 & 70)[132] and hamamelitannin (95)[195] have also been shown to be effective in inhibiting biofilm formation. Hamameltannin inhibits the QS regulator RNAIII and shown to reduce biofilm formation in MRSA and methicillin-resistant S. epidermidis strains [195]. Other small molecules that reduce biofilm include baicalein (97), which is a natural product shown to inhibit P. aeruginosa biofilm formation at 20 µM [199]. Baicalein promoted proteolysis of TraR protein in E. coli, strongly suggesting that its biofilm-inhibition properties might be due to QS inhibition [199]. With the exception of few biofilm inhibitors including AHL analogs (63, 64, 69 & 70) and the brominated furanones (79–82), whose targets are known, the majority of biofilm inhibitor that have been discovered so far have unknown targets. These include 2-aminoimidazole/triazole (98)[200,201], several structurally diverse small molecules capable of inhibiting Pseudomonas aeruginosa biofilm formation (99–104; Figure 16)[202], SB2 (105) and ureidothiophene (106)[203]. 2-aminoimidazole/triazole (98) is a simple derivative of the marine natural product bromoageliferin, which can disperse established proteobacterial biofilm as well as inhibit biofilm formation [200,201]. 98 was able to disperse S. aureus biofilm with an IC50 of 2.6 µM. The mechanism of biofilm formation or dispersion remains to be elucidated. (Z)-2-decenoic acid (107) is also a potent biofilm-dispersion inducer, being able to completely disperse P. aeruginosa biofilm at a concentration of 2.5 nM [204].

    In an attempt to increase the pool of small molecules that can inhibit biofilm formation, the Clardy group developed an efficient high-throughput screen to identify small-molecules that are effective against P. aeruginosa biofilm development [202]. A total of 66,095 compounds were screened via a luminescence-based biofilm assay for both attachment and detachment of bacteria to solid surfaces. Several small molecules belonging to different structural classes were identified as effective biofilm formation inhibitors see Table 4.

    SB2 (105) and ureidothiophene (106) are antibiofilm molecules that inhibit transcription in vitro against S. epidermidis[203]. It has been suggested that their antibiofilm activities may be connected with their action as RNA polymerase inhibitors but this claim has yet to be validated [203].

    In an unexpected twist, recent studies have shown that known drugs such as azithromycin [205], ceftazidime and tobramycin [206] inhibit biofilm formation at subminimal inhibitory concentration. It remains to be seen if the antibiofilm activities of these established drugs are derived from their known targets, 23S rRNA (tobramycin), 23S rRNA (azithromycin) and bacterial cell wall (ceftazidine).

    ▪ New targets of antibiotic therapy

    Recent efforts on antibiotic discovery have focused on new bacterial target proteins that have important roles in viability. Several RNA polymerase inhibitors including lipiarmycin (116; Figure 17) are currently in clinical trials [203]. Other new antibiotic targets include LpxC, an enzyme that is involved in the synthesis of lipopolysaccharide, Lipid A [207,208]. A few inhibitors of bacterial lipid biosynthesis are in clinical trials including Rs-DPLa (110), eritoran (117) and CHIR-090 (111; Figure 17)[207,208]. Type II fatty-acid synthesis proteins FabF and FabH are also important antibiotic targets. Platensimycin (108)[209] and platencin (109)[210], novel antibiotics produced by Streptomyces platensis, are nanomolar FabF inhibitor and FabF/H dual inhibitor, respectively. Simplified analogs thereof are expected to become useful next generation antibiotics [211–214]. However, a recent paper has questioned the validity of using type II fatty-acid synthesis inhibitors as effective antibiotics [215]. The FtsZ protein, which is homologous to tubulin and is important for prokaryotic cell division [216], has recently been shown to be a drugable target. It has been demonstrated that compound 113(Figure 17) can inhibit FtsZ protein leading to bacterial cell death [216,217]. RecA is an essential enzyme required in bacterial cells for the repair and maintenance of DNA [218]. The small molecule, suramin (112) has been shown to inhibit RecA’s ATPase activity [219]. Thiopeptide antibiotics, such as thiostrepton (119; Figure 17) and mirococcin, have also drawn the attention of medicinal chemists and biochemists. These thiopeptide antibiotics bind to the L11 domain of 23S ribosomal RNA [220] but, because of their poor solubility, none of the antibiotics in this family have found clinical use. The synthesis of soluble analogs of thiostreptin (119) may allow clinical application in the future [221,222]. Similarly lantibiotics, a class of antibiotics that are ribosomally synthesized and contain the unusually lanthionine amino acid residue, have bacteriocidal effects. This is due to the ability of lantibiotics to inhibit peptidoglycan formation [223]. Examples of lantibiotics include nisin (118; Figure 17) and epidermin, along with many others. Recent interest in the lantibiotics is due to the fact that they have been used in food preservation for decades without any observed resistance emerging. Nisin (118) is not currently in clinical use owing to the fact that it is degraded in the human digestive system [224]. The development of analogs, which are stable in the human intestine and gut, could allow these compounds to have clinical use in the future [225,226]. Inhibition of peptide deformylase is another novel strategy for antibiotic therapy, owing to its importance in protein synthesis [227]. The natural product actinonin (115) has been identified as an inhibitor of peptide deformylase [227] and other synthetic compounds that inhibit deformylase have been designed and are in clinical trials. Furthermore, inhibition of teichoic acid biosynthesis has been targeted through the inhibition of LtaS, the lipoteichoic acid synthase. Teichoic acid is a major component of Gram-positive bacteria cell walls. Teichoic acid attached directly to the cell wall (wall teichoic acid) or linked to membrane lipids (lipoteichoic acid) have been shown to be vital for bacterial survival.

    ▪ Novel c-di-GMP signaling presents new targets for novel antibiotic developments

    In the coming decades, as more information from bacterial genome sequencing becomes available, it is anticipated that new proteins that are involved in bacterial regulatory pathways will be unveiled. The exact details of transduction processes such as c-di-GMP signaling will also be uncovered. c-di-GMP (Figure 18) has emerged as an important intracellular signaling molecule uniquely present in bacteria [228].

    Since first being discovered as a cellulose synthase activator in Gluconacetobacter xylinus (formally known as Acetobactor xylinum) by Benziman over two decades ago [229], c-di-GMP has been shown to regulate diverse cellular processes, including the expression of virulence factors and genes that are important for the formation and maintenance of biofilm [51]. In addition, some advances have now been made regarding the elucidation of the roles of QS and c-di-GMP signaling in biofilm formation [230].

    As c-di-GMP signaling and QS regulate the same complex processes, including biofilm formation and virulence, it has been argued that these two signaling pathways may be linked [149]. Accumulating evidence shows that c-di-GMP and QS signaling intersect [231]. For example in V. cholerae, QS autoinducers control the expression levels of small regulatory RNAs known as Qrr sRNAs [232]. These sRNAs are involved in an Hfq-mediated degradation of target mRNAs such as HapR (a master transcriptional regulator). Recently, the Bassler’s laboratory has shown that Qrr sRNA also disrupts an inhibitory stem loop structure of vca0939 mRNA [233]. This mRNA translates into an enzyme that forms c-di-GMP. Additionally, in V. cholerae a transcription factor known as AphA, which is regulated by QS, influences the expression of DGCs and PDEAs[234]. Recently, Wood showed that TpbA, a tyrosine phosphatase that is controlled by LasR, regulates biofilm formation by inhibiting intracellular c-di-GMP concentrations [230]. The exact details of c-di-GMP signaling is currently being studied by several laboratories and it is expected that analogs of c-di-GMP [235] or other small molecules that will inhibit the proteins that are involved in the biosynthesis of c-di-GMP will become useful as either antivirulence or antibiofilm drugs.

    Future perspective

    Appreciation of the amazing ability of bacteria to evolve mechanisms that render bacteriocidal or bacteriostatic antibiotics ineffective has led to a re-evaluation of how to develop new antibiotics and manage bacterial infections. In the last decade our understanding of signaling processes in bacteria (QS and c-di-GMP signaling, which leads to bacterial virulence factor production and biofilm formation) has increased. This development, coupled with the recent discovery that all bacteriocidal antibiotics lead to production of hydroxyl radical and oxidative damage (Figure 19)[236], has rekindled a belief that the antibiotic resistance problem might not lead to a plague in the future. Judging by the flurry of activities in the antibiotic research field, there is every hope that the new paradigm shift in antibiotic discovery will result in effective next-generation antibiotics. In the coming years, it is anticipated that several receptors for the universal QS molecule, AI-2, will be identified and biochemically characterized. Additionally, the structure of AI-3 is expected to be solved and more structural data for Lux-R type receptors that bind AHLs are also expected to be obtained. Furthermore, the adaptor proteins and RNA molecules that bind to c-di-GMP and transmit the binding event into virulence factors expression or biofilm formation are more likely to be identified and biochemically characterized. This projected wealth of information will give the medicinal chemist a powerful platform for developing the next generation antibiotics that aim to curb bacterial virulence instead of killing bacteria. Most of the anti-QS agents reported are experimental, and comprehensive pharmacokinetic data on these molecules are lacking. More anti-QS drugs that do not have in vivo instability issues and are drug like are desired. The demonstration that AI-1 and analogs thereof have immunological profiles [237] calls for a full scientific investigation of the immune suppressive activities of autoinducer analogs using relevant disease models. It is very plausible that a decade from now, bacterial infections will be treated with combination therapeutics consisting of multiple chemical entities that will function as antibiofilm, antivirulence, bacteriocidal or bacteriostatic and antiresistance. The real challenge will be finding a cocktail of such drugs that do not suffer from unwanted drug–drug interactions.

    Table 1.  Classical antibiotics.
    AntibioticsStructural classYear introduced into US marketTarget processYear resistance observedRef.
    Prontosil (1)Sulfonamides1935Inhibition of folate synthesis (binds dihydropteroate synthase)1940s[10]
    Penicillin G (2)β-lactams1943Inhibition of cell wall synthesis (binds transpeptidases required for peptidoglycan synthesis)1940s[11]
    Streptomycin (8)Aminoglycosides1943Inhibition of protein synthesis (binds 16S rRNA of 30S subunit at tRNA acceptor A site)1959[4]
    Chloramphenicol (9)Phenylpropanoid/ amphenicols1947Inhibition of protein synthesis (binds 23S rRNA of 50S subunit at P site)1959[5]
    Tetracycline (6)Polyketides1948Inhibition of protein synthesis (binds 16S rRNA at 30S ribosomal subunit at the A site inhibits ribosome interaction with aminoacyl-tRNA)1953[6]
    Erythromycin (7)Macrolides1952Inhibition of protein synthesis (binds 23S rRNA of 50S subunit at the P site; interferes with peptide chain elongation)1982[7]
    Vancomycin (14)Glycopeptides1956Inhibition of cell wall synthesis (binds D-Ala-D-Ala terminus; inhibits peptidoglycan synthesis)1984[8]
    Polymyxin B (17)Polypeptides1960sDisruption of outer and inner membranes (binds lipopolysaccharide)1976[12]
    Methicillin (3)β-lactams1960Inhibition of cell wall synthesis1961[3]
    Ciprofloxacin (13)Quinolone1987Inhibition of DNA replication (binds DNA gyrase or topoisomerase IV)1988[13]
    Cephalothin (4)Cephalosporins1964Inhibition of cell wall synthesis1965[14]
    Rifampicin (12)Rifamycins1967Inhibition of DNA-dependent RNA polymerase (binds β subunit)1968[15]
    Clindamycin (16)Lincosamides1969Inhibition of protein synthesis (binds 23S rRNA of 50S subunit at P site)1979[16]
    Imipenem (5)Carbapenem1985Inhibition of cell wall synthesis (binds transpeptidases)1986[17]
    Quinupristin/ dalfopristin (11)Streptogramin1999Inhibition of protein synthesis (binds 23S rRNA of 50S subunit at the P site)1999[18]
    Linezolid (10)Oxazolidinones2000Inhibition of protein synthesis (binds 23S rRNA of 50S subunit at P site)2001[19]
    Daptomycin (15)Lipopeptides2003Depolarization of bacterial membrane2005[20]

    Other quinolones were discovered in 1962[2].

    Other streptogramin antibiotics were developed prior to this year including virginiamycin and pristinamycin.

    Data from[1,2]

    Table 2.  New antibiotics approved between 2004 and 2009.
    Antibiotic (trade name)Drug/structural class (lead compound)Year introduced into US marketTarget processIndication
    Telithromycin (18) (Ketek®)Ketolides (erythromycin)2004Inhibition of protein synthesis (binds domain II of 50S ribosome)Bronchitis, sinusitis and pneumonia
    Rifaximin (19) (Xifaxan®)Rifamycins (rifampicin)2004Inhibition of DNA-dependent RNA synthesisTravelers’ diarrhea
    Tigecycline (23) (Tygacil®)Polyketides (tetracycline)2005Inhibition of protein synthesis (binds 30S ribosome)Skin and soft tissue infections; effective against MRSA and Acinetobacter baumannii
    Doripenem (20) (Finibax®/Doribax®)Carbapenem (imipenem)2007Inhibition of cell wall synthesis (targets penicillin binding proteins)Intra-abdominal and urinary tract infections
    Azithromycin (Azasite®)Macrolides (erythromycin)2007Inhibition of protein synthesis (binds 50S ribosome)Conjunctivitis
    Retapamulin (21) (Altabax®/Altargo®)Pleuromutilin2007Inhibition of protein synthesis (binds domain V of 50S ribosome)Impetigo due to S. aureus or S. pyogenes
    Amoxicillin (Moxatag®)β-lactams (penicillin)2008Inhibition of cell wall synthesisTonsillitis and pharyngitis
    Besifloxacin (22) (Besivance®)Quinolones (ciprofloxacin)2009Inhibition of bacterial DNA gyrase and topisomeraseBacterial conjunctivitis; Gram-negative and Gram-positive
    Telavancin (24) (Vibativ®)Lipoglycopeptides (vancomycin)2009Inhibition of cell wall synthesis and depolarization of bacterial membraneSkin infections; effective against MRSA and VRE

    These antibiotics were already in the market but were reintroduced for new indications.

    MRSA: Methicillin-resistant Staphylococcus aureus; VRE: Vancomycin-resistant enterococcus.

    Data taken from[45–47].

    Table 3.  Quorum-sensing inhibitors.
    Small moleculeKiProteinSource of proteinRef.
    570.37 µMLuxSBacillus subtilis[118]
    BuT-DADMe-ImmA (58a)296 fMMTANEscherichia coli[114]
    pClPhT-DADMe-ImmA (58b)47 fMMTANE. coli[114]
     IC50Agr groupOrganism tested 
    trAIP-I D2A (62)5 nMAgrC-1Staphylococcus aureus[121]
    AIP-IV Y5F (59)2 nMAgrC-2S. aureus[121]
    AIP-I D5A (60)0.3 nMAgrC-3S. aureus[121]
    AIP-I D5N (61)20 nMAgrC-4S. aureus[121]
      Transcription factorReporter assay 
    p-bromophenyl-AHL (63)3.89 µMLasRE. coli DH5α (pJN105L pSC11)[123]
    Indole- AHL (64)0.83 µMLasRE. coli DH5α (pJN105L pSC11)[123]
    Cyclohexyl-AHL (65)2.69 µM§LuxRVibrio fischeri ES114[123]
    C4-sulfonyl-AHL (66)0.61 µMTraRAgrobacterium tumefaciens WCF47 (pCF372)[123]
    PD12 (73)30 nM#LasRPseudomonas aeruginosa MW1 (pUM15)[149]

    Inhibition assay was conducted in the presence of the group-specific wild-type AIP agonist at 100 nM.

    Tested against 7.5 nM odDHL.

    §Tested against 5 µM OHHL.

    Tested against 100 nM OOHL.

    #Tested against 0.3 µM 3OC12-HSL.

    Table 4.  Antibiofilm activity of select small molecules.
    Biofilm inhibitorInhibitory concentration (µM)OrganismRef.
    Baicalein (97)20Pseudomonas aeruginosa[199]
    2-AIT (98)0.53P. aeruginosa (PA14)[200]
    (99)0.53P. aeruginosa[202]
    (100)1.55P. aeruginosa[202]
    (101)1.77P. aeruginosa[202]
    (102)2.52P. aeruginosa[202]
    (103)2.87P. aeruginosa[202]
    (104)3.85P. aeruginosa[202]
    SB2 (105)2.50§Staphylococcus epidermidis[203]
    Ureidothiophene (106)0.50§S. epidermidis[203]
    (Z)-2-decenoic acid (107)0.0025P. aeruginosa (PAO1)[204]

    Compound98has an IC50 of 0.81 µM against S. aureus; IC50 values of 0.98, 5.6 and 9.5 µM against A. baumannii, P.aeruginosa (PAO1) and B. bronchiseptica (RB50), respectively.

    These values are inhibitory concentrations and were reported as EC50 values.

    §These are IC50 values.

    Compound107was shown to induce dispersion with an efficacy of 24.6%[204].

    Antibiotic resistance

    Ability of a microorganism to withstand the effects of compounds aimed to kill or inhibit its growth

    Biofilm

    Aggregation of micro-organisms on living and nonliving surfaces that are resistant to antibiotic treatment

    Quorum sensing

    Form of cell-to-cell communication whereby the secretion and detection of small molecules allow bacteria to coordinate the expression of genes

    Autoinducer-1

    Acylhomoserine lactones used as intra-species signaling molecules by Gram-negative bacterium

    Autoinducer-2

    Family of cyclic furanones derived from 4,5-dihydroxy-2,3-pentadione and synthesized by LuxS, found in both Gram-negative and Gram-positive bacteria

    Autoinducer-3

    Unidentified aromatic compound used in inter-kingdom signaling

    Virulence factors

    Molecules or proteins expressed and secreted by pathogens and which are harmful to host cells

    Pseudomonas aeruginosa

    Gram-negative, aerobic bacterium; an opportunistic human pathogen, known to cause lung infections in cystic fibrosis patients

    Staphylococcus aureus

    Gram-positive, facultatively anaerobic, bacterium that is often found as a commensal on human skin but also associated with food poisoning, skin and other infections

    Vibrio cholerae

    Gram-negative, facultatively anaerobic, bacterium that causes acute diarrheal illness known as cholera

    Salmonella typhimurium

    Gram-negative bacterium that causes gastroenteritis in humans. It is commonly found in the intestinal lumen

    Escherichia coli

    Gramnegative bacterium commonly found in the lower intestine of humans

    Diguanylate cyclase (DGC)

    Enzyme that makes cyclic diguanylic acid from two molcules of GTP

    Phosphodiesterase A (PDEA)

    Enzyme that degrades cyclic diguanylic acid into the linear dimer pGpG

    Executive summary

    • ▪ Bacteria have shown in the past that almost any essential gene is evolvable. Bacterial strains that have resistance mechanisms for all current antibiotics exist.

    • ▪ A paradigm shift in treating bacterial infections is to stop killing bacteria as this creates evolutionary pressure to evolve resistance mechanisms. Instead, virulence factor production must be curtailed.

    • ▪ Two signaling pathways, quorum sensing and cyclic diguanylic acid signaling, are master regulators of bacterial virulence factors production and biofilm formation.

    • ▪ Quorum sensing and cyclic diguanylic acid signaling in particular are not completely understood.

    • ▪ Small molecules that can inhibit quorum sensing in pathogenic bacteria have been identified.

    • ▪ Interestingly, effective molecules that can also sequester or inhibit bacterial toxins such as shiga or cholera toxins have been developed.

    • ▪ Future efforts will unravel intricate details and more protein/RNA targets in both quorum sensing and cyclic diguanylic acid signaling. However, the real challenge will be the development of drug-like molecules that will find clinical use.

    Acknowledgements

    The authors thank Deana Jaber for proof reading and help with some of the compounds in Figure 16. We also thank the reviewers of this manuscript for excellent suggestions.

    Financial & competing interests disclosure

    This work was supported by the University of Maryland and National Science Foundation grant CHE0746446. Jacqueline AI Smith is a recipient of Ministry of Education GANN fellowship. The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

    No writing assistance was utilized in the production of this manuscript.

    Papers of special note have been highlighted as: ▪ of interest ▪▪ of considerable interest

    Bibliography

    • Arias CA, Murray BE. Antibiotic-resistant bugs in the 21st century – a clinical super challenge. N. Engl. J. Med.360(5),439–443 (2009).
    • von Nussbaum F, Brands M, Hinzen B, Weigand S, Häbich D. Antibacterial natural products in medicinal chemistry – exodus or revival? Angew. Chem. Int. Ed.45(31),5072–5129 (2006).
    • Jevons MP. ‘Celbenin’-resistant Staphylococci. Br. Med. J.1(5219),124–125 (1961).
    • Watanabe T, Watanabe M. Transduction of streptomycin resistance in Salmonella typhimurium. J. Gen. Microbiol.21,16–29 (1959).
    • Ramsey CH, Edwards PR. Resistance of Salmonellae isolated in 1959 and 1960 to tetracyclines and chloramphenicol. Appl. Microbiol.9(5),389–391 (1961).
    • Franklin TJ, Godfrey A. Resistance of Escherichia coli to tetracyclines. Biochem. J.94,54–60 (1965).
    • Haddy RI, Gordon RC, Shamiyeh L, Wofford R, Fechner L, Sahanek E. Erythromycin resistance in group A β-hemolytic Streptococci. Pediatr. Infect. Dis.1(4),236–238 (1982).
    • Shlaes DM, Marino J, Jacobs MR. Infection caused by vancomycin-resistant Streptococcus sanguis II. Antimicrob. Agents Chemother.25(4),527–528 (1984).
    • Uttley AH, Collins CH, Naidoo J, George RC. Vancomycin-resistant enterococci. Lancet331(8575–8576),57–58 (1988).
    • 10  Sevag MG, Green MN. The mechanism of resistance to sulfonamides. I. Factors controlling the formation of arylamine from tryptophane by Staphylococcus aureus. J. Bacteriol.48(6),615–622 (1944).
    • 11  Barber M, Whitehead JE. Bacteriophage types in penicillin-resistant staphylococcal infection. Br. Med. J.2(4627),565–569 (1949).
    • 12  Gilleland HE Jr, Murray RG. Ultrastructural study of polymyxin-resistant isolates of Pseudomonas aeruginosa. J. Bacteriol.125(1),267–281 (1976).
    • 13  Kaatz GW, Seo SM. Mechanism of ciprofloxacin resistance in Pseudomonas aeruginosa. J. Infect. Dis.158(3),537–541 (1988).
    • 14  Benner EJ, Bennett JV, Brodie JL, Kirby WM. Inactivation of cephalothin and cephaloridine by Staphylococcus aureus. J. Bacteriol.90(6),1599–1604 (1965).
    • 15  Ezekiel DH, Hutchins JE. Mutations affecting RNA polymerase associated with rifampicin resistance in Escherichia coli. Nature220(5164),276–277 (1968).
    • 16  Tally FP, Snydman DR, Gorbach SL, Malamy MH. Plasmid-mediated, transferable resistance to clindamycin and erythromycin in Bacteroides fragilis. J. Infect. Dis.139(1),83–88 (1979).
    • 17  Cuchural GJ Jr, Malamy MH, Tally FP. β-lactamase-mediated imipenem resistance in Bacteroides fragilis. Antimicrob. Agents Chemother.30(5),645–648 (1986).
    • 18  Schmitz FJ, Sadurski R, Stattfeld A, Kray A, Verhoef J, Fluit AC. Cross-resistance analyses and molecular typing of Staphylococcus aureus and Streptococcus spp. isolates resistant to quinupristin/dalfopristin. J. Antimicrob. Chemother.44(6),847–849 (1999).
    • 19  Tsiodras S, Gold HS, Sakoulas G et al. Linezolid resistance in a clinical isolate of Staphylococcus aureus. Lancet358(9277),207–208 (2001).
    • 20  Sabol K, Patterson JE, Lewis JS II, Owens A, Cadena J, Jorgensen JH. Emergence of daptomycin resistance in Enterococcus faecium during daptomycin therapy. Antimicrob. Agents Chemother.49(4),1664–1665 (2005).
    • 21  Sabath LD, Wheeler N, Laverdiere M, Blazevic D, Wilkinson BJ. A new type of penicillin resistance of Staphylococcus aureus. Lancet1(8009),443–447 (1977).
    • 22  Fisher JF, Meroueh SO, Mobashery S. Bacterial resistance to β-lactam antibiotics. compelling opportunism, compelling opportunity. Chem. Rev.105(2),395–424 (2005).
    • 23  Finlay J, Miller L, Poupard JA. A review of the antimicrobial activity of clavulanate. Antimicrob. Chemother.52(1),18–23 (2003).
    • 24  Imtiaz U, Billings EM, Knox JR, Mobashery S. A structure-based analysis of the inhibition of class A β-lactamases by sulbactam. Biochemistry33(19),5728–5738 (1994).
    • 25  Rasheed JK, Jay C, Metchock B et al. Evolution of extended-spectrum β-lactam resistance (SHV-8) in a strain of Escherichia coli during multiple episodes of bacteremia. Antimicrob. Agents Chemother.41(3),647–653 (1997).
    • 26  Shaw KJ, Rather PN, Hare RS, Miller GH. Molecular genetics of aminoglycoside resistance genes and familial relationships of the aminoglycoside-modifying enzymes. Microbiol. Mol. Biol. Rev.57(1),138–163 (1993).
    • 27  Gonzales RD, Schreckenberger PC, Graham MB, Kelkar S, DenBesten K, Quinn JP. Infections due to vancomycin-resistant Enterococcus faecium resistant to linezolid. Lancet357(9263). 1179 (2001).
    • 28  Hiramatsu K, Hanaki H, Ino T, Yabuta K, Oguri T, Tenover FC. Methicillin-resistant Staphylococus aureus clinical strain with reduced vancomycin susceptibility. J. Antimicrob. Chemother.40(1),135–136 (1997).
    • 29  Cetinkaya Y, Falk PF, Mayhall CG. Vancomycin-resistant enterococci. Clin. Microbiol. Rev.13(4),686–707 (2000).
    • 30  Sulavik MC, Houseweart C, Cramer C et al. Antibiotic susceptibility profiles of Escherichia coli strains lacking multidrug efflux pump genes. Antimicrob. Agents Chemother.45(4),1126–1136 (2001).
    • 31  Poole K, Krebes K, McNally C, Neshat S. Multiple antibiotic resistance in Pseudomonas aeruginosa. evidence for involvement of an efflux operon. J. Bacteriol.175(22),7363–7372 (1993).
    • 32  Ziha-Zarifi I, Llanes C, Köhler T, Pechere JC, Plesiat P. In vivo emergence of multidrug-resistant mutants of Pseudomonas aeruginosa overexpressing the active efflux system MexA-MexB-OprM. Antimicrob. Agents Chemother.43(2),287–291 (1999).
    • 33  Kaatz GW, Seo SM. Mechanisms of fluoroquinolone resistance in genetically related strains of Staphylococcus aureus. Antimicrob. Agents Chemother.41(12),2733–2737 (1997).
    • 34  Piddock LJ. Multidrug-resistance efflux pumps-not just for resistance. Nat. Rev. Microbiol.4(8),629–636 (2006).
    • 35  Hsieh PC, Siegel SA, Rogers B, Davis D, Lewis K. Bacteria lacking a multidrug pump. a sensitive tool for drug discovery. Proc. Natl Acad. Sci. USA95(12),6602–6606 (1998).
    • 36  Mullin S, Mani N, Grossman TH. Inhibition of antibiotic efflux in bacteria by the novel multidrug resistance inhibitors biricodar (VX-710) and timcodar (VX-853). Antimicrob. Agents Chemother.48(11),4171–4176 (2004).
    • 37  Pumbwe L, Chang A, Smith RL, Wexler HM. Clinical significance of overexpression of multiple RND-family efflux pumps in Bacteroides fragilis isolates. J. Antimicrob. Chemother.58(3),543–548 (2006).
    • 38  Miyamae S, Ueda O, Yoshimura F, Hwang J, Tanaka Y, Nikaido H. A MATE family multidrug efflux transporter pumps out fluoroquinolones in Bacteroides thetaiotaomicron. Antimicrob. Agents Chemother.45(12),3341–3346 (2001).
    • 39  Debono M, Abbott BJ, Molloy RM et al. Enzymatic and chemical modifications of lipopeptide antibiotic A21978C. the synthesis and evaluation of daptomycin (LY146032). J. Antibiot. (Tokyo)41(8),1093–1105 (1988).
    • 40  Tally FP, DeBruin MF. Development of daptomycin for gram-positive infections. J. Antimicrob. Chemother.46(4),523–526 (2000).
    • 41  Brickner SJ, Hutchinson DK, Barbachyn MR et al. Synthesis and antibacterial activity of U-100592 and U-100766, two oxazolidinone antibacterial agents for the potential treatment of multidrug-resistant gram-positive bacterial infections. J. Med. Chem.39(3),673–679 (1996).
    • 42  Barbachyn MR, Ford CW. Oxazolidinone structure-activity relationships leading to linezolid. Angew. Chem. Int. Ed.42(18),2010–2023 (2003).
    • 43  Jones RN, Della-Latta P, Lee LV, Biedenbach DJ. Linezolid-resistant Enterococcus faecium isolated from a patient without prior exposure to an oxazolidinone. Report from the SENTRY Antimicrobial Surveillance Program. Diagn. Microbiol. Infect. Dis.42(2),137–139 (2002).
    • 44  Hayden MK, Rezai K, Hayes RA, Lolans K, Quinn JP, Weinstein RA. Development of Daptomycin resistance in vivo in methicillin-resistant Staphylococcus aureus. J. Clin. Microbiol.43(10),5285–5287 (2005).
    • 45  Owens J. 2006 drug approvals. finding the niche. Nat. Rev. Drug Discov.6(2),99–101 (2007).
    • 46  Hughes B. 2007 FDA drug approvals: a year of flux. Nat. Rev. Drug Discov.7(2),107–109 (2008).
    • 47  Hughes B. 2008 FDA drug approvals. Nat. Rev. Drug Discov.8(2),93–96 (2009).
    • 48  Yan K, Madden L, Choudhry AE, Voigt CS, Copeland RA, Gontarek RR. Biochemical characterization of the interactions of the novel pleuromutilin derivative retapamulin with bacterial ribosomes. Antimicrob. Agents Chemother.50(11),3875–3881 (2006).
    • 49  Givskov M, Hentzer M. Jamming bacterial communication: new strategies to combat bacterial infections and the development of biofilm. In: Bacterial Cell-to-Cell Communication: Role in Virulence and Pathogenesis. Demuth DR, Lamont RJ (Eds.). Cambridge University Press, Cambridge, UK, 65–90 (2006).
    • 50  Kulesekara H, Lee V, Brencic A et al. Analysis of Pseudomonas aeruginosa diguanylate cyclases and phosphodiesterases reveals a role for bis-(3´-5´)-cyclic-GMP in virulence. Proc. Natl Acad. Sci. USA103(8),2839–2844 (2006).
    • 51  Jenal U. Cyclic di-guanosine-monophosphate comes of age: a novel secondary messenger involved in modulating cell surface structures in bacteria? Curr. Opin. Microbiol.7(2),185–191 (2004).
    • 52  Chan WC, Coyle BJ, Williams P. Virulence regulation and quorum sensing in staphylococcal infections: competitive AgrC antagonists as quorum sensing inhibitors. J. Med. Chem.47(19),4633–4641 (2004).▪ Reviews natural autoinducing peptides and discusses the possibility that synthetic peptides or peptoids could be developed as antivirulence agents.
    • 53  Ji G, Beavis R, Novick RP. Bacterial interference caused by autoinducing peptide variants. Science276(5321),2027–2030 (1997).
    • 54  Wright JS III, Jin R, Novick RP. Transient interference with staphylococcal quorum sensing blocks abscess formation. Proc. Natl Acad. Sci. USA102(5),1691–1696 (2005).
    • 55  Whiteley M, Lee KM, Greenberg EP. Identification of genes controlled by quorum sensing in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA96(24) 13904–13909 (1999).
    • 56  Pearson JP, Gray KM, Passador L et al. Structure of the autoinducer required for expression of Pseudomonas aeruginosa virulence genes. Proc. Natl Acad. Sci. USA91(1),197–201 (1994).
    • 57  Smith RS, Iglewski BH. P. aeruginosa quorum-sensing systems and virulence. Curr. Opin. Microbiol.6(1),56–60 (2003).
    • 58  Ahmer BM, van Reeuwijk J, Timmers CD, Valentine PJ, Heffron F. Salmonella typhimurium encodes an SdiA homolog, a putative quorum sensor of the LuxR family, that regulates genes on the virulence plasmid. J. Bacteriol.180(5),1185–1193 (1998).
    • 59  Michael B, Smith JN, Swift S, Heffron F, Ahmer BMM. SdiA of Salmonella enterica is a LuxR homolog that detects mixed microbial communities. J.Bacteriol.183(19) 5733–5742 (2001).
    • 60  Janssens JC, Metzger K, Daniels R et al. Synthesis of N-acyl homoserine lactone analogues reveals strong activators of SdiA, the Salmonella enterica serovar Typhimurium LuxR homologue. Appl. Environ. Microbiol.73(2),535–544 (2007).
    • 61  Cvitkovitch DG, Li YH, Ellen RP. Quorum sensing and biofilm formation in Streptococcal infections. J. Clin. Invest.112(11),1626–1632 (2003).
    • 62  Davies DG, Parsek MR, Pearson JP, Iglewski BH, Costerton JW, Greenberg EP. The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science280(5361),295–298 (1998).
    • 63  Singh PK, Schaefer AL, Parsek MR, Moninger TO, Welsh MJ, Greenberg EP. Quorum-sensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature407(6805),762–764 (2000).
    • 64  Rickard AH, Palmer RJ Jr, Blehert DS et al. Autoinducer 2: a concentration-dependent signal for mutualistic bacterial biofilm growth. Mol. Microbiol.60(6),1446–1456 (2006).
    • 65  Ahmed NA, Petersen FC, Scheie AA. AI-2 quorum sensing affects antibiotic susceptibility in Streptococcus anginosus. J. Antimicrob. Chemother.60(1),49–53 (2007).
    • 66  Waters CM, Bassler BL. Quorum sensing: cell-to-cell communication in bacteria. Ann. Rev. Cell Dev. Biol.21,319–346 (2005).
    • 67  Federle MJ, Bassler BL. Interspecies communication in bacteria. J. Clin. Invest.112(9),1291–1299 (2003).
    • 68  Hughes DT, Sperandio V. Inter-kingdom signaling: communication between bacteria and their hosts. Nat. Rev. Microbiol.6(2),111–120 (2008).
    • 69  Walters M, Sircili MP, Sperandio V. AI-3 synthesis is not dependent on luxS in Escherichia coli. J. Bacteriol.188(16),5668–5681 (2006).
    • 70  Cámara M, Hardman A, Williams P, Milton D. Quorum sensing in Vibrio cholerae. Nat. Genet.32(2),217–218 (2002).
    • 71  Higgins DA, Pomianek ME, Kraml CM, Taylor RK, Semmelhack MF, Bassler BL. The major Vibrio cholerae autoinducer and its role in virulence factor production. Nature450(7171),883–886 (2007).
    • 72  Clarke MB, Sperandio V. LuxS-dependent regulation of Escherichia coli virulence. In: Bacterial Cell-to-Cell Communication: Role in Virulence and Pathogenesis. Demuth DR, Lamont RJ (Eds). Cambridge University Press, Cambridge, UK, 159–168 (2006).
    • 73  Nakayama J, Cao Y, Horii T et al. Gelatinase biosynthesis-activating pheromone: a peptide lactone that mediates a quorum sensing in Enterococcus faecalis. Mol. Microbiol.41(1),145–154 (2001).
    • 74  Geske GD, O’Neill JC, Blackwell HE. Expanding dialogues: from natural autoinducers to non-natural analogues that modulate quorum sensing in Gram-negative bacteria. Chem. Soc. Rev.37,1432–1447 (2008).
    • 75  Horinouchi S, Ohnishi Y, Kang DK. The A factor regulatory cascade and cAMP in the regulation of physiological and morphological development in Streptomycesgriseus. J. Ind. Microbiol. Biotechnol.27,177–182 (2001).
    • 76  Pesci EC, Milbank JB, Pearson JP et al. Quinolone signaling in the cell-to-cell communication system of Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA96(20),11229–11234 (1999).
    • 77  Vendeville A, Winzer K, Heurlier K, Tang CM, Hardie KR. Making ‘sense’ of metabolism: autoinducer-2, LuxS and pathogenic bacteria. Nat. Rev. Microbiol.3(5),383–396 (2005).
    • 78  Xavier KB, Bassler BL. LuxS quorum sensing: more than just a numbers game. Curr. Opin. Microbiol.6(2),191–197 (2003).
    • 79  Chen X, Schauder S, Potier N et al. Structural identification of a bacterial quorum-sensing signal containing boron. Nature415(6871),545–549 (2002).▪▪ Revealed the identity of the universal quorum-sensing autoinducer AI-2.
    • 80  Miller ST, Xavier KB, Campagna SR et al.Salmonella typhimurium recognizes a chemically distinct form of the bacterial quorum-sensing signal AI-2. Mol. Cell15(5),677–687 (2004).
    • 81  Xue T, Zhao L, Sun H, Zhou, X, Sun B. LsrR-binding site recognition and regulatory characteristics in Escherichia coli AI-2 quorum sensing. Cell Res.19(11),1258–1268 (2009).
    • 82  Kim SY, Lee SE, Kim YR et al. Regulation of Vibrio vulnificus virulence by the LuxS quorum-sensing system. Mol. Microbiol.48(6),1647–1664 (2003).
    • 83  Shin NR, Lee DY, Shin SJ, Kim KS, Yoo HS. Regulation of proinflammatory mediator production in RAW264.7 macrophage by Vibrio vulnificus LuxS and SmcR. FEMS Immunol. Med. Microbiol.41,169–176 (2004).
    • 84  Coulthurst SJ, Leopold Kurz C, Salmond GPC. luxS mutants of Serratia defective in autoinducer-2-dependent ‘quorum sensing’ show strain-dependent impacts on virulence and production of carbapenem and prodigiosin. Microbiol.150,1901–1910 (2004).
    • 85  Ohtani K, Hayashi H, Shimizu T. The luxS gene is involved in cell-cell signalling for toxin production in Clostridium perfringens. Mol. Microbiol.44(1),171–179 (2002).
    • 86  Winzer K, Hardie KR, Burgess N et al. LuxS. its role in central metabolism and the in vitro synthesis of 4-hydroxy-5-methyl-3-(2H)-furanone. Microbiol.148(Pt 4),909–922 (2002).
    • 87  Tavender TJ, Halliday NM, Hardie KR Winzer K. LuxS-independent formation of AI-2 from ribulose-5-phosphate. BMC Microbiol.8(98),2188–2198 (2008).
    • 88  Hauck T, Hübner Y, Brühlmann F, Schwab W. Alternative pathway for the formation of 4,5-dihydroxy-2,3-pentanedione, the proposed precursor of 4-hydroxy-5-methyl-3(2H)-furanone as well as autoinducer-2, and its detection as natural constituent of tomato fruit. Biochim. Biophys. Acta.1623(2–3),109–119 (2003).
    • 89  Li J, Wang L, Hashimoto Y et al. A stochastic model of Escherichia coli AI-2 quorum signal circuit reveals alternative synthesis pathways. Mol. Syst. Biol.2,67 (2006).
    • 90  Pesci EC, Pearson JP, Seed PC, Iglewski BH. Regulation of las and rhl quorum sensing in Pseudomonas aeruginosa. J. Bacteriol.179(10),3127–3132 (1997).
    • 91  Chugani SA, Whiteley M, Lee KM, D’Argenio D, Manoil C, Greenberg EP. QscR, a modulator of quorum-sensing signal synthesis and virulence in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA98(5),2752–2757 (2001).
    • 92  Ledgham F, Ventre I, Soscia C, Foglino M, Sturgis JN, Lazdunski A. Interaction of the quorum sensing regulator QscR: interation with itself and the other regulators of Pseudomonas aeruginosa LasR and RhlR. Mol. Microbiol.48(1),199–210 (2003).
    • 93  Wagner VE, Bushnell D, Passador L, Brooks AI, Iglewski BH. Microarry analysis of Pseudomonas aeruginosa quorum-sensing regulons: effects of growth phase and environment. J. Bacteriol.185(7),2080–2095 (2003).
    • 94  Schuster M, Lostroh CP, Ogi T, Greenberg EP. Identification, timing, and signal specificity of Pseudomonas aeruginosa quorum-controlled genes: a transcriptome analysis. J. Bacteriol.185(7),2066–2079 (2003).
    • 95  Smith RS, Iglewski BH. Pseudomonas aeruginosa quorum sensing as a potential antimicrobial target. J. Clin. Invest.112(10),1460–1465 (2003).
    • 96  Gambello MJ, Iglewski BH. Cloning and characterization of the Pseudomonas aeruginosa lasR gene, a transcriptional activator of elastase expression. J. Bacteriol.173(9),3000–3009 (1991).
    • 97  Kessler E, Safrin M, Gustin JK, Ohman DE. Elastase and the LasA protease of Pseudomonas aeruginosa are secreted with their propeptides. J. Biol. Chem.273(46),30225–30231 (1998).
    • 98  Gambello MJ, Kaye S, Iglewski BH. LasR of Pseudomonas aeruginosa is a transcriptional activator of the alkaline protease gene (apr) and an enhancer of exotoxin A expression. Infect. Immun.61(4),1180–1184 (1993).
    • 99  Chapon-Hervé V, Akrim M, Latifi A, Williams P, Lazdunski A, Bally M. Regulation of the xcp secretion pathway by multiple quorum-sensing modulons in Pseudomonas aeruginosa. Mol. Microbiol.24(6),1169–1178 (1997).
    • 100  Holden MTG, Chhabra SR, de Nys R et al. Quorum-sensing cross talks: isolation and chemical characterization of cyclic dipeptides from Pseudomonas aeruginosa and other gram-negative bacteria. Mol. Microbiol.33(6),1254–1266 (1999).
    • 101  Campbell J, Lin Q, Geske GD, Blackwell HE. New and unexpected insights into the modulation of LuxR-type quorum sensing by cyclic dipeptides. ACS Chem. Biol.4(12),1051–1059 (2009).
    • 102  Novick RP. Autoinduction and signal transduction in the regulation of staphylococcal virulence. Mol. Microbiol.48(6),1429–1449 (2003).
    • 103  Yarwood JM. Quorum-sensing-dependent regulation of staphylococcal virulence and biofilm development. In: Bacterial Cell-to-Cell Communication: Role in Virulence and Pathogenesis. Demuth DR, Lamont RJ (Eds). Cambridge University Press, Cambridge, UK, 199–266 (2006).
    • 104  Cheung AL, Zhang G. Global regulation of virulence determinants in Staphylococcus aureus by the SarA protein family. Front. Biosci.7,d1825–1842 (2002).
    • 105  Xavier KB, Miller ST, Lu W et al. Phosphorylation and processing of the quorum-sensing molecule autoinducer-2 in enteric bacteria. ACS Chem. Biol.2(2) 128–136 (2007).
    • 106  Novick RP, Jiang D. The staphylococcal saeRS system coordinates environmental signals with agr quorum sensing. Microbiology149(Pt 10),2709–2717 (2003).
    • 107  Lee MS, Morrison DA. Identification of a new regulator in Streptococcus pneumoniae linking quorum sensing to competence for genetic transformation. J. Bacteriol.181(16),5004–5016 (1999).
    • 108  Lazazzera BA, Solomon JM, Grossman AD. An exported peptide functions intracellularly to contribute to cell density signaling in B. subtilis. Cell89(6),917–925 (1997).
    • 109  Kelly RC, Bolitho ME, Higgins DA et al. The Vibrio cholerae quorum-sensing autoinducer CAI-1:analysis of the biosynthetic enzyme CqsA. Nat. Chem. Biol.5(12),891–895 (2009).
    • 110  Wang L, Hashimoto Y, Tsao CY, Valdes JJ, Bentley WE. Cyclic AMP (cAMP) and cAMP receptor protein influence both synthesis and uptake of extracellular autoinducer 2 in Escherichia coli. J. Bacteriol.187(6),2066–2076 (2005).
    • 111  Moré MI, Finger LD, Stryker JL, Fuqua C, Eberhard A, Winans SC. Enzymatic synthesis of a quorum-sensing autoinducer through use of defined substrates. Science272(5268),1655–1658 (1996).
    • 112  Schauder S, Shokat K, Surette MG, Bassler BL. The LuxS family of bacterial autoinducers: biosynthesis of a novel quorum-sensing signal molecule. Mol. Microbiol.41(2) 463–476 (2001).
    • 113  Musk DJ Jr, Hergenrother PJ. Chemical coutermeasures for the control of bacterial biofilms: effective compounds and promising targets. Curr. Med. Chem.13(18),2163–2177 (2006).
    • 114  Singh V, Evans GB, Lenz DH et al. Femtomolar transition state analogue inhibitors of 5´-methylthioadenosine/ S-adenosylhomocysteine nucleosidase from Escherichia coli. J. Biol. Chem.280(18),18265–18273 (2005).▪ Demonstrated successful inhibition of quorum sensing in bacteria via the inhibition of autoinducer synthesis.
    • 115  Gutierrez JA, Crowder T, Rinaldo-Matthis A, Ho MC, Almo SC, Schramm VL. Transition state analogs of 5´ methylthioadenosine nucleosidase discrupt quorum sensing. Nature Chem. Biol.5(4),251–257 (2009).
    • 116  Gopishetty B, Zhu J, Rajan R et al. Probing the catalytic mechanism of S-ribosylhomocysteinase (LuxS) with catalytic intermediates and substrate analogues. J. Am. Chem. Soc.131(3),1243–1250 (2009).
    • 117  Alfaro JF, Zhang T, Wynn DP, Karschner EL, Zhou ZS. Synthesis of LuxS inhibitors targeting bacterial cell-cell communication. Org. Lett.6(18),3043–3046 (2004).
    • 118  Shen G, Rajan R, Zhu J, Bell CE, Pei D. Design and synthesis of substrate and intermediate analogue inhibitors of S-ribosylhomocysteinase. J. Med. Chem.49(10),3003–3011 (2006).
    • 119  Wnuk SF, Roert J, Sobczak AJ et al. Inhibition of S-ribosylhomocysteinase (LuxS) by substrate analogues modified at the ribosyl C-3 position. Bioorg. Med. Chem.17(18),6699–6706 (2009).
    • 120  Han X, Lu C. Biological activity and identification of a peptide inhibitor of LuxS from Streptococcus suis serotype 2. FEMS Microbiol. Lett.294(1),16–23 (2009).
    • 121  Lyon GJ, Wright JS, Muir TW, Novick RP. Key determinants of receptor activation in the agr autoinducing peptides of Staphylococcus aureus. Biochemistry41(31),10095–10104 (2002).
    • 122  Scott RJ, Lian LY, Muharram SH et al. Side-chain-to-tail thiolactone peptide inhibitors of the staphylococcal quorum-sensing system. Bioorg. Med. Chem. Lett.13(15),2449–2453 (2003).
    • 123  Geske GD, O’Neill JC, Miller DM, Mattmann ME, Blackwell HE. Modulation of bacterial quorum sensing with synthetic ligands: systematic evaluation of N-acylated homoserine lactones in multiple species and new insights into their mechanisms of action. J. Am. Chem. Soc.129(44),13613–13625 (2007).
    • 124  Smith KM, Bu Y, Suga H. Induction and inhibition of Pseudomonas aeruginosa quorum sensing by synthetic autoinducer analogs. Chem. Biol.10,81–89 (2003).
    • 125  Smith KM, Bu Y, Suga H. Library screening for synthetic agonists and antagonists of a Pseudomonas aeruginosa autoinducer. Chem. Biol.10(6),563–571 (2003).
    • 126  Persson T, Givskov M, Nielsen J. Quorum sensing inhibition: targeting chemical communication in Gram-negative bacteria. Curr. Med. Chem.12(26),3103–3115 (2005).
    • 127  Swern LR, Swem DL, O’Loughlin CT et al. A quorum-sensing antagoist targets both membrane-bound and cytoplasmic receptors and controls bacterial pathogenicity. Mol. Cell.35(2),143–153 (2009).
    • 128  Ni N, Li M, Wang J, Wang B. Inhibitors and antagonists of bacterial quorum sensing. Med. Res. Rev.29(1),66–124 (2009).
    • 129  Lowery CA, Park J, Kaufmann GF, Janda KD. An unexpected switch in the modulation of AI-2-based quorum sensing discovered through synthetic 4,5-dihydroxy-2,3-pentanedione analogues. J. Am. Chem. Soc.130(29),9200–9201 (2008).
    • 130  Ganin H, Tang X, Meijler MM. Inhibiton of Pseudomonas aeruginosa quorum sensing by AI-2 analogs. Bioorg. Med. Chem. Lett.19(14),3941–3944 (2009).
    • 131  Smith JAI, Wang J, Nguyen-Mau SM, Lee V, Sintim HO. Biological screening of a diverse set of AI-2 analogs in Vibrio harveyi suggests that receptors which are involved in synergistic agonism of AI-2 and analogues are promiscuous. Chem. Commun.7(35),7033–7035 (2009).
    • 132  Geske GD, Wezeman RJ, Siegel AP, Blackwell HE. Small molecule inhibitors of bacterial quorum sensing and biofilm formation. J. Am. Chem. Soc.127(37),12762–12763 (2005).
    • 133  Williams SC, Patterson EK, Carty NL, Griswold JA, Hamood AN, Rumbaugh KP. Pseudomonas aeruginosa autoinducer enters and functions in mammalian cells. J. Bacteriol.186(8),2281–2287 (2004).
    • 134  Smith RS, Fedyk ER, Springer TA, Mukaida N, Iglewski BH, Phipps RP. IL-8 production in human lung fibroblasts and epithelial cells activated by the Pseudomonas autoinducer N-3-oxododecanoyl homoserine lactone is transcriptionally regulated by NF-kB and activator protein-2. J. Immunol.167,366–374 (2001).
    • 135  Schuster M, Greenberg EP. LuxR-type proteins in Pseudomonas aeruginosa quorum sensing: distinct mechanisms with global implications. In: Cell–Cell Communication in Bacteria. Winans SC, Bassler BL (Eds). ASM Press, Washington, DC, USA, 133–144 (2008).
    • 136  Schuster M, Urbanowski ML, Greenberg EP. Promoter specificity in Pseudomonas aeruginosa quorum sensing revealed by DNA binding of purified LasR. Proc. Natl Acad. Sci. USA101(45),15833–15839 (2004).
    • 137  Weingart CL, White CE, Liu S et al. Direct binding of the quorum sensing regulator CepR of Burkhoderia cenocepacia to two target promoters in vitro. Mol. Microbiol.57(2),452–467 (2005).
    • 138  Urbanowski ML, Lostroh CP, Greenberg EP. Reversibile acyl-homoserine lactone binding to purified Vibrio fischeri LuxR protein. J. Bacteriol.186(3),631–637 (2004).
    • 139  Lee JH, Lequette Y, Greenberg EP. Activity of purified QscR, a Pseudomonas aeruginosa orphan quorum-sensing transcription factor. Mol. Microbiol.59(2),602–609 (2006).
    • 140  Medina G, Juárez K, Valderrama B, Soberón-Chávez G. Mechanism of Pseudomonas aeruginosa RhlR transcriptional regulation of the rhlAB promoter. J. Bacteriol.185(20),5976–5983 (2003).
    • 141  Castang S, Reverchon S, Gouet P, Nasser W. Direct evidence for the modulation of the activity of the Erwinia chrysanthemi quorum-sensing regulator ExpR by acylhomoserine lactone pheromone. J. Biol. Chem.281(40),29972–29987 (2006).
    • 142  Wagner VE, Iglewski BH. Quorum sensing and regulation of Pseudomonas aeruginosa infections. In: Bacterial Cell-to-Cell Communication: Role in Virulence and Pathogenesis. Demuth DR, Lamont RJ (Eds). Cambridge University Press, Cambridge, UK, 1–15 (2006).
    • 143  Lewenza S, Conway B, Greenberg EP, Sokol PA. Quorum sensing in Burkholderia cenocepacia: identification of the LuxRI homologs CepRI. J. Bacteriol.181(3),748–756 (1999).
    • 144  Aharoni A, Gaidukov L, Yagur S et al. Directed evolution of mammalian paraoxonases PON1 and PON3 for bacterial expression and catalytic specialization. Proc. Natl Acad. Sci. USA101(2),482–487 (2004).
    • 145  Kharsonsky O, Tawfik DS. Structure–reactivity studies of serum paraoxonase PON1 suggest that its native activity is lactonase. Biochem.44(16),6371–6382 (2005).
    • 146  Greenberg EP, Chun CK, Ozer EA, Welsh MJ, Zabner J. Inactivation of a Pseudomonas aeruginosa quorum-sensing signal by human airway epithelia. Proc. Natl Acad. Sci. USA101(10),3587–3590 (2004).▪▪ Demonstrated that human hosts have evolved strategies to quench bacterial quorum sensing. This paper also cautions against the development of synthetic antiquorum sensing analogs that are based on the lactone core.
    • 147  Yang F, Wang L-H, Wang J et al. Quorum quenching enzyme activity is widely conserved in the sera of mammalian species. FEBS Lett.579(17),3713–3717 (2005).
    • 148  Jakubowski H. Calcium-dependent human serum homocysteine thiolactone hydrolase. A protective mechanism against protein N-homocysteinylation. J. Biol. Chem.275(6),3957–3962 (2000).
    • 149  Müh U, Schuster M, Heim R, Singh A, Olson ER, Greenberg EP. Novel Pseudomonas aeruginosa quorum-sensing inhibitors identified in an ultra-high-throughput screen. Antimicrob. Agent Chemother.50(11),3674–3679 (2006).
    • 150  Persson T, Hansen TH, Rasmussen TB, Skindersø ME, Givskov M, Nielsen J. Rational design and synthesis of new quorum-sensing inhibitors derived from acylated homoserine lactones and natural products from garlic. Org. Biomol. Chem.3(2),253–262 (2005).
    • 151  Rasko DA, Moreira CG, Li DR et al. Targeting QseC signaling and virulence for antibiotic development. Science321(5892),1078–1080 (2008).
    • 152  Janssens JC, Steenackers H, Robijns S et al. Brominated furanones inhibit biofilm formation by Salmonella enterica serovar typhimurium. Appl. Environ. Microbiol.74(21),6639–6648 (2008).
    • 153  Hentzer M, Riedel K, Rasmussen TB et al. Inhibition of quorum sensing in Pseudomonas aeruginosa biofilm bacteria by a halogenated furanone compound. Microbiology148(Pt 1),87–102 (2002).
    • 154  Wu H, Song Z, Hentzer M et al. Synthetic furanones inhibit quorum sensing and enhance bacterial clearance in Pseudomonas aeruginosa lung infection in mice. J. Antimicrob. Chemother.53(6),1054–1061 (2004).
    • 155  Defoirdt T, Miyamoto CM, Wood TK et al. The natural furanone (5Z)-4-bromo-5-(bromomethylene)-3-butyl-2(5H)-furanone disrupts quorum sensing-regulated gene expression in Vibrio harveyi by decreasing the DNA-binding activity of the transcriptional regulator protein luxR. Environ. Microbiol.9(10),2486–2495 (2007).
    • 156  Zang T, Lee BWK, Cannon LM et al. A naturally occurring brominated furanone covalently modifies and inactivates LuxS. Bioorg. Med. Chem. Lett.19(21),6200–6204 (2009).
    • 157  Rasmussen TB, Bjarnsholt T, Skindersoe ME et al. Screening for quorum-sensing inhibitors (QSI) by use of a novel genetic system, the QSI selector. J. Bacteriol.187(5),1799–1814 (2005).
    • 158  Rasmussen TB, Skindersoe ME, Bjarnsholt T et al. Identity and effects of quorum-sensing inhibitors produced by Penicillium species. Microbiology151(Pt 5),1325–1340 (2005).
    • 159  Skindersoe ME, Ettinger-Epstein P, Rasmussen TB et al. Quorum sensing antagonism from marine organisms. Marin. Biotech.10(1),56–63 (2008).
    • 160  Brackman G, Defoirdt T, Miyamoto C et al. Cinnamaldehyde and cinnamaldehye derivatives reduce virulence in Vibrio spp. by decreasing the DNA-binding activity of the quorum sensing response regulator LuxR. BMC Microbiol. DOI: 10.1186/1471-2180-2188-2149 (2008) (Epub ahead of print).
    • 161  De Lamo Marin S, Xu Y, Meiljer MM, Janda KD. Antibody catalyzed hydrolysis of a quorum sensing signal found in Gram-negative bacteria. Bioorg. Med. Chem. Lett.17(6),1549–1552 (2007).
    • 162  Kapadnis PB, Hall E, Ramstedt M et al. Towards quorum-quenching catalytic antibodies. Chem. Comm.538–540 (2009).
    • 163  Kaufmann GF, Sartorio R, Lee SH et al. Antibody interference with N-acyl homoserine lactone-mediated bacterial quorum sensing. J. Am. Chem. Soc.128(9),2802–2803 (2006).
    • 164  Park J, Jagasia R, Kaufmann GF et al. Infection control by antibody disruption of bacterial quorum sensing signaling. Chem. Biol.14(10),1119–1127 (2007).
    • 165  Miyairi S, Tateda K, Fuse ET et al. Immunization with 3-oxododecanoyl-L-homoserine lactone-protein conjugate protects mice from lethal Pseudomonas aeruginosa lung infection. J. Med. Microbiol.55(Pt 10),1381–1387 (2006).
    • 166  Sun D, Accavitti MA, Bryers JD. Inhibition of biofilm formation by monoclonal antibodies against Staphylococcus epidermidis RP62A accumulation-associated protein. Clin. Diagn. Lab. Immunol.12(1),93–100 (2005).
    • 167  Sio CF, Otten LG, Cool RH et al. Quorum quenching by an N-acyl-homoserine lactone acylase from Pseudomonas aeruginosa PAO1. Infect. Immun.74(3),1673–1682 (2006).
    • 168  Dong YH, Wang LH, Xu JL, Zhang HB, Zhang XF, Zhang LH. Quenching quorum-sensing-dependent bacterial infection by an N-acyl homoserine lactonase. Nature411(6839),813–817 (2001).
    • 169  Roy V, Fernandes R, Tsao C-Y, Bentley WE. Cross species quorum quenching using a native AI-2 processing enzyme. ACS Chem. Bio. DOI: 10.1021/cb9002738 (2009) (Epub ahead of print).
    • 170  Ukkonen P, Varis K, Jernfors M et al. Treatment of acute otitis media with an antiadhesive oligosaccharide: a randomised, double-blind, placebo-controlled trial. Lancet356(9239),1398–1402 (2000).
    • 171  Kunz C, Rodriguez-Palmero M, Koletzko B, Jensen R. Nutritional and biochemical properties of human milk, Part I: general aspects, proteins, and carbohydrates. Clin. Perinatol.26(2),307–333 (1999).
    • 172  Newburg DS. Are all human milks created equal? Variation in human milk oligosaccharides. J. Pediatr. Gastroenterol. Nutr.30(2),131–133 (2000).
    • 173  Basha S, Rai P, Poon V et al. Polyvalent inhibitors of anthrax toxin that target host receptors. Proc. Natl Acad. Sci. USA103(36),13509–13513 (2006).▪▪ Identified an effective molecule that blocks anthrax toxin entry into host cell. By targeting the host cell’s receptors and not the bacterial toxin, resistance towards the molecule is diminished.
    • 174  Firon N, Ofek I, Sharon N. Carbohydrate specificity of the surface lectins of Escherichia coli, Klebsiella pneumoniae and Salmonella typhimurium. Carbohydr. Res.120,235–249 (1983).
    • 175  Mysore JV, Wigginton T, Simon PM, Zopf D, Heman-Ackah LM, Dubois A. Treatment of Helicobacter pylori infection in rhesus monkeys using a novel anti-adhesion compound. Gastroenterol.117(6),1316–1325 (1999).
    • 176  Foo LY, Lu Y, Howell AB, Vorsa N. A-type proanthocyanidin trimers from cranberry that inhibit adherence of uropathogenic P-fimbriated Escherichia coli. J. Nat. Prod.63(9),1225–1228 (2000).
    • 177  Suree N, Yi SW, Thieu W et al. Discovery and structure-activity relationship analysis of Staphylococcus aureus sortase A inhibitors. Bioorg. Med. Chem.17(20),7174–7185 (2009).
    • 178  Pinkner JS, Remaut H, Buelens F et al. Rationally designed small compounds inhibit pilus biogenesis in uropathogenic bacteria. Proc. Natl Acad. Sci. USA103(47),17897–17902 (2006).
    • 179  Aberg V, Almqvist F. Pilicides-small molecules targeting bacterial virulence Org. Biomol. Chem.5(12),1827–1834 (2007).
    • 180  Tseng TT, Tyler BM, Setubal JC. Protein secretion systems in bacterial-host associations, and their description in the gene ontology. BMC Microbiol.9(Suppl. 1),S2 (2009).
    • 181  Sandkvist M. Type II secretion and pathogenesis. Infect. Immun.69(6),3523–3535 (2001).
    • 182  Galán JE, Wolf-Watz H. Protein delivery into eukaryotic cells by type III secretion machines. Nature444(7119),567–573 (2006).
    • 183  Cascales E, Christie PJ. The versatile bacterial type IV secretion systems. Nat. Rev. Microbiol.1(2),137–149 (2003).
    • 184  Henderson IR, Navarro-Garcia F, Desvaux M, Fernandez RC, Ala’Aldeen D. Type V protein secretion pathway: the autotransporter story. Microbiol. Mol. Biol. Rev.68(4),692–744 (2004).
    • 185  Pukatzki S, Ma AT, Sturtevant D et al. Identification of a conserved bacterial protein secretion system in Vibrio cholerae using the Dictyostelium host model system. Proc. Natl Acad. Sci. USA103(5),1528–1533 (2006).
    • 186  Cascales E. The Type VI secretion toolkit. EMBO Rep.9(8),735–741 (2008).
    • 187  Sperandio V, Mellies JL, Nguyen W, Shin S, Kaper JB. Quorum sensing controls expression of the type III secretion gene transcription and protein secretion in enterohemorrhagic and enteropathogenic Escherichia coli. Proc. Natl Acad. Sci. USA96(26) 15196–15201 (1999).
    • 188  Gauthier A, Robertson ML, Lowden M, Ibarra JA, Puente JL, Finlay BB. Transcriptional inhibitor of virulence factors in enteropathogenic Escherichia coli. Antimicrob. Agents Chemother.49(10),4101–4109 (2005)
    • 189  Henke JM, Bassler BL. Quorum sensing regulates Type III secretion in Vibrio harveyi and Vibrio parahaemolyticusJ. Bacteriol.186(12),3794–3805 (2004).
    • 190  Kauppi AM, Nordfelth R, Uvell H, Wolf-Watz H, Elofsson M. Targeting bacterial virulence: inhibitors of type III secretion in Yersinia. Chem. Biol.10(3),241–249 (2003).
    • 191  Hung DT, Shakhnovich EA, Pierson E, Mekalanos JJ. Small-molecule inhibitor of Vibrio cholerae virulence and intestinal colonization. Science310(5748),670–674 (2005).
    • 192  Kitov PI, Sadowska JM, Mulvey G et al. Shiga-like toxins are neutralized by tailored multivalent carbohydrate ligands. Nature403(6770),669–672 (2000).
    • 193  Simpson LS, Burdine L, Dutta AK, Feranchak AP, Kodadek T. Selective toxin sequestrants for the treatment of bacterial infection. J. Am. Chem. Soc.131(16),5760–5762 (2009).
    • 194  Merritt EA, Hol WG. AB5 toxins. Curr. Opin. Struct. Biol.5(2),165–171 (1995).
    • 195  Kiran MD, Adikesavan NV, Cirioni O et al. Discovery of a quorum-sensing inhibitor of drug-resistant staphylococcal infections by structure-based virtual screening. Mol. Pharmacol.73(5),1578–1586 (2008).
    • 196  Peters L, König GM, Wright AD et al. Secondary metabolites of Flustra foliacea and their influence on bacteria. Appl. Environ. Microbiol.69(6),3469–3475 (2003).
    • 197  Böttcher T, Sieber SA. β-lactones as specific inhibitors of ClpP attenuate the production of extracellular virulence factors of Staphylococcus aureus. J. Am. Chem. Soc.130(44),14400–14401 (2008).
    • 198  Frees D, Qazi SN, Hill PJ, Ingmer H. Alternative roles of ClpX and ClpP in Staphylococcus aureus stress tolerance and virulence. Mol. Microbiol.48(6),1565–1578 (2003).
    • 199  Zeng Z, Qian L, Cao L et al. Virtual screening for novel quorum sensing inhibitors to eradicate biofilm formation of Pseudomonas aeruginosa. Appl. Microbiol. Biotechnol.79(1),119–126 (2008).
    • 200  Rogers SA, Melander C. Construction and screening of a 2-aminoimidazole library identifies a small molecule capable of inhibiting and dispersing bacterial biofilms across order, class, and phylum. Angew. Chem. Int. Ed.47,5229–5231 (2008).
    • 201  Richards JJ, Huigens III RW, Ballard TE, Basso A, Cavanagh J, Melander, C. Inhibition and dispersion of proteobacterial biofilms. Chem. Commun.14(14) 1698–1700 (2008).
    • 202  Junker LM, Clardy J. High-throughput screens for small-molecule inhibitors of Pseudomonas aeruginosa biofilm development. Antimicrob. Agents Chemother.51(10),3582–3590 (2007).
    • 203  Villain-Guillot P, Gualtieri M, Bastide L, Leonetti, JP. In vitro activities of different inhibitors of bacterial transcription against Staphylococcus epidermidis biofilm. Antimicrob. Agents Chemother.51(9),3117–3121 (2007).
    • 204  Davies DG, Marques CNH. A fatty acid messenger is responsible for inducing dispersion in microbial biofilms. J. Bacteriol.191(5),1393–1403 (2009).
    • 205  Nalca Y, Jänsch L, Bredenbruch F, Geffers R, Buer J, Häussler S. Quorum-sensing antagonistic activities of azithromycin in Pseudomonas aeruginosa PAO1: a global approach. Antimicrob. Agents Chemother.50(5),1680–1688 (2006).
    • 206  Garske LA, Beatson SA, Leech AJ, Walsh SL, Bell SC. Sub-inhibitory concentrations of ceftazidime and tobramycin reduce the quorum sensing signals of Pseudomonas aeruginosa. Pathology36(6),571–575 (2004).
    • 207  Barb AW, Jiang L, Raetz CR, Zhou P. Structure of the deactylase LpxC bound to the antibiotic CHIR-090: time-dependent inhibition and specificity in ligand binding. Proc. Natl Acad. Sci. USA104(47),18433–18438 (2007).
    • 208  Barb AW, McClerren AL, Snehelatha K, Reynolds CM, Zhou P, Raetz CR. Inhibition of lipid A biosynthesis as the primary mechanism of CHIR-090 antibiotic activity in Escherichia coli. Biochemistry46(12),3793–3802 (2007).
    • 209  Wang J, Soisson SM, Young K et al. Platensimycin is a selective FabF inhibitor with potent antibiotic properties. Nature441(7091),358–361 (2006).
    • 210  Wang J, Kodali S, Lee SH et al. Discovery of platencin, a dual FabF and FabH inhibitor with in vivo antibiotic properties. Proc. Natl Acad. Sci. USA104(18),7612–7616 (2007).
    • 211  McNulty J, Nair JJ, Capretta A. A synthesis of sulfonamide analogs of platensimycin employing a palladium-mediated carbonylation strategy. Tetrahedron Lett.50(28),4087–4091 (2009).
    • 212  Nicolaou KC, Tria GS, Edmonds DJ, Kar M. Total synthese of (±)-platencin and (-)-platencin. J. Am. Chem. Soc.131(43),15909–15917 (2009).
    • 213  Barykina OV, Rossi KL, Rybak MJ, Snider BB. Synthesis and antibacterial properties of (-)-nor-platencin. Org. Lett.11(22),5334–5337 (2009).
    • 214  Wang J, Lee V, Sintim HO. Efforts towards the identification of simpler platensimycin analogues-the total synthesis of oxazinidinyl platensimycin. Chem. Eur. J.15(12),2747–2750 (2009).
    • 215  Brinster S, Lamberet G, Staels B, Trieu-Cuot P, Gruss A, Poyart C. Type II fatty acid synthesis is not a suitable antibiotic target for Gram-positive pathogens. Nature458(7234),83–86 (2009).
    • 216  de Boer P, Crossley R, Rothfield L. The essential bacterial cell-division protein FtsZ is a GTPase. Nature359(6392),254–256 (1992).
    • 217  Dasgupta D. Novel compound with potential of an antibacterial drug targets FtsZ protein. Biochem. J.423(1),E1–E3 (2009).
    • 218  Lusetti SL, Cox MM. The bacterial RecA protein and the recombinational DNA repair of stalled replication forks. Ann. Rev. Biochem.71,71–100 (2002).
    • 219  Wigle TJ, Singleton SF. Directed molecular screening for RecA ATPase inhibitors. Bioorg. Med. Chem. Lett.17(12),3249–3253 (2007).
    • 220  Lentzen G, Klinck R, Matassova N, Aboul-ela F, Murchie AIH. Structural basis for contrasting activities of ribosome binding thiazole antibiotics. Chem. Biol.10(8),769–778 (2003).
    • 221  Nicolaou KC, Nevalainen M, Zak M, Bulat S, Bella M, Safina BS. Synthetic studies on thiostrepton: construction of thiostrepton analogues with the thiazoline-containing macrocycle. Angew. Chem. Int. Ed.42(29),3418–3424 (2003).
    • 222  Kelly WL, Pan L, Li C. Thiostrepton biosynthesis: prototype for a new family of bacteriocins. J. Am. Chem. Soc.131(12),4327–4334 (2009).
    • 223  Chatterjee C, Paul M, Xie L, van der Donk WA. Biosynthesis and mode of action of lantibiotics. Chem. Rev.105(2),633–684 (2005).
    • 224  Delves-Broughton J, Blackburn P, Evans RJ, Hugenholtz J. Applications of the bacteriocin, nisin. Anton. Van Leeuwen.69(2),193–202 (1996).
    • 225  Bregant S, Tabor AB. Orthogonally protected lanthionines: synthesis and use for the solid-phase synthesis of an analogues of nisin ring C. J. Org. Chem.70(7),2430–2438 (2005).
    • 226  Li B, Yu JP, Brunzelle JS, Moll GN, van der Donk WA, Nair SK. Structure and mechanism of the lantibiotic cyclase involved in nisin biosynthesis. Science311(5766),1464–1467 (2006).
    • 227  Chen DZ, Patel DV, Hackbarth CJ et al. Actinonin, a naturally occurring antibacterial agent, is a potent deformylase inhibitor. Biochemistry39(6),1256–1262 (2000).
    • 228  Römling U, Gomelsky M, Galperin MY. C-di-GMP: the dawning of a novel bacterial signalling system. Mol. Microbiol.57(3),629–639 (2005).
    • 229  Ross P, Weinhouse H, Aloni Y et al. Regulation of cellulose synthesis in Acetobacter xylinum by cyclic diguanylic acid. Nature325(6101),279–281 (1987).
    • 230  Ueda A, Wood TK. Connecting quorum sensing, c-di-GMP, pel polysaccharide, and biofilm formation in Pseudomonas aeruginosa through tyrosine phosphatase TpbA (PA3885). PloS Pathogens5(6),e1000483 (2009).▪ Showed a clear link between quorum sensing and cyclic diguanylic acid signaling.
    • 231  Waters CM, Lu W, Rabinowitz JD, Bassler BL. Quorum sensing controls biofilm formation in Vibrio cholerae through modulation of cyclic di-GMP levels and repression of vpsT. J. Bacteriol.190(7),2527–2536 (2008).
    • 232  Lenz DH, Mok KC, Lilley BN, Kulkarni RV, Wingreen NS, Bassler BL. The small RNA chaperone Hfq and multiple small RNAs control quorum sensing in Vibrio harveyi and Vibrio cholerae. Cell118(1),69–82 (2004).
    • 233  Hammer BK, Bassler BL. Regulatory small RNAs circumvent the conventional quorum sensing pathway in pandemic Vibrio cholerae. Proc. Natl Acad. Sci. USA104(27),11145–11149 (2007).
    • 234  Kovacikova G, Lin W, Skorupski K. Dual regulation of genes involved in acetoin biosynthesis and motility/ biofilm formation by the virulence activator AphA and the acetate-responsive LysR-type regulator AlsR in Vibrio cholerae. Mol. Microbiol.57(2),420–433 (2005).
    • 235  Kiburu I, Shurer A, Yan L, Sintim HO. A simple solid-phase synthesis of the ubiquitous bacterial signaling molecule, c-di-GMP and analogues. Mol. Biosyst.4(6),518–520 (2008).
    • 236  Kohanski MA, Dwyer DJ, Hayete B, Lawrence CA, Collins JJ. A common mechanism of cellular death induced by bactericidal antibiotics. Cell130(5),797–810 (2007).
    • 237  Chhabra SR, Harty C, Hooi DS et al. Synthetic analogues of the bacterial signal (quorum sensing) moleculhe N-(3-oxododecanoyl)-L-homoserine lactone as immune modulators. J. Med. Chem.46(1),97–104 (2003).