Next Article in Journal
Chitosan-Integrated Curcumin–Graphene Oxide/Copper Oxide Hybrid Nanocomposites for Antibacterial and Cytotoxicity Applications
Previous Article in Journal
Culex quinquefasciatus Mosquitoes Resist Acquisition of Methicillin-Resistant Staphylococcus aureus: Insights from Field and Laboratory Studies
Previous Article in Special Issue
Essential Oils from Southern Italian Aromatic Plants Synergize with Antibiotics against Escherichia coli, Pseudomonas aeruginosa and Enterococcus faecalis Cell Growth and Biofilm Formation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Quorum Quenching Approaches against Bacterial-Biofilm-Induced Antibiotic Resistance

by
Patrizia D’Aquila
1,
Elisabetta De Rose
1,
Giada Sena
1,
Angelo Scorza
2,
Bonaventura Cretella
2,
Giuseppe Passarino
1 and
Dina Bellizzi
1,*
1
Department of Biology, Ecology and Earth Sciences, University of Calabria, 87036 Rende, Italy
2
Villa Ermelinda, Progetto Terza Età, 88842 Cutro, Italy
*
Author to whom correspondence should be addressed.
Antibiotics 2024, 13(7), 619; https://doi.org/10.3390/antibiotics13070619
Submission received: 10 June 2024 / Revised: 28 June 2024 / Accepted: 1 July 2024 / Published: 3 July 2024
(This article belongs to the Special Issue Molecular Mechanism of Antibiotic Resistance in Microbial Biofilms)

Abstract

:
With the widespread phenomenon of antibiotic resistance and the diffusion of multiple drug-resistant bacterial strains, enormous efforts are being conducted to identify suitable alternative agents against pathogenic microorganisms. Since an association between biofilm formation and antibiotic resistance phenotype has been observed, a promising strategy pursued in recent years focuses on controlling and preventing this formation by targeting and inhibiting the Quorum Sensing (QS) system, whose central role in biofilm has been extensively demonstrated. Therefore, the research and development of Quorum Quenching (QQ) compounds, which inhibit QS, has gradually attracted the attention of researchers and has become a new strategy for controlling harmful microorganisms. Among these, a number of both natural and synthetic compounds have been progressively identified as able to interrupt the intercellular communication within a microbial community and the adhesion to a surface, thus disintegrating mature/preformed biofilms. This review describes the role played by QS in the formation of bacterial biofilms and then focuses on the mechanisms of different natural and synthetic QS inhibitors (QSIs) exhibiting promising antibiofilm ability against Gram-positive and Gram-negative bacterial pathogens and on their applications as biocontrol strategies in various fields.

1. Introduction

Quorum Sensing (QS) is a global gene regulation mechanism based on bacterial cell-to-cell communication achieved through the release, detection, and response of extracellular signaling molecules called autoinducers (AIs) [1]. It can be considered an adaptive response to the high cellular density at which the cumulative production of AIs occurs, enabling detection and response through a specific transcriptional regulator. In Gram-positive and Gram-negative bacteria, different types of QS systems have been widely described, and different classes of AIs have been identified [2]. QS allows bacteria to perceive their surroundings, regulate their density and behavior, and optimize the use of available nutrients, giving them the ability to live as multicellular organisms. What is more, it makes the cell community manage self-competition as well as collectively interact with their host. QS regulates several biological processes including bioluminescence, sporulation, competence, antibiotic production, and virulence factor production by pathogenic bacteria, as well as the formation and maintenance of biofilm, the complex, three-dimensional structures representing a long-established survival mechanism for bacteria [3]. The high cell density that is created in the biofilm makes it the ideal condition for QS regulation. In fact, numerous QS-dependent genes code for the biosynthesis of exopolysaccharides that directly participate in the architecture of mature biofilms. Biofilms play a fundamental role in pathogenesis as they can evade the human immune response, and bacteria forming them exhibit resistance to treatment with various antibiotics, making fighting these infections even more challenging, leading to prolonged illness and contributing significantly to high morbidity and mortality rates [4]. The important role played by biofilms in microbial infections is more than relevant if we consider that the National Institutes of Health (NIH) revealed that among all microbial and chronic infections, 65% and 80%, respectively, are associated with biofilm formation. These infections are contracted in hospital settings by patients undergoing long-term hospitalization or in close contact with medical devices and implants on which biofilms form, increasing the risk of infection during surgeries or other medical procedures. They are colonized with extreme efficiency by microorganisms of the skin flora, including Gram-positive bacteria such as Staphylococcus aureus, which use them as “entry routes” to the patient’s internal organs, causing infections with often very serious outcomes [5].
Starting from the above consideration, recently, control strategies have been developed to activate so-called Quorum Quenching (QQ), thus providing new possibilities for dealing with infectious diseases and overcoming and solving the problem of microbial resistance. QQ interferes specifically with the QS system of bacteria, hindering the exchange of information between them and reducing the expression level of hazard factors. It may be mainly achieved by inhibiting the production of AI molecules, by inhibiting AI detection through the inactivation of the receptors, or by enzyme-catalyzed degradation or modification of the AI molecules. Different classes of QQ enzymes as well as QQ inhibitors have been identified so far, thus highlighting how they exhibit substrate specificities [6].
The present paper aims to review the current knowledge regarding the implication of the biofilm process in the phenomenon of antibiotic resistance and the QQ strategies. The first section describes the stages of biofilm formation and its structural elements as well as its intrinsic characteristics that are the basis of biofilm-induced antibiotic resistance. The next section reviews the mechanisms of inhibition of the QS signaling pathway for controlling biofilm formation and then focuses on the antibiofilm mechanisms of different natural and synthetic QS inhibitors. The final section reports the strategies which are currently being adopted as biocontrol solutions in various fields.

2. Biofilm and Antibiotic Resistance

Biofilm is an aggregation of microbial cells attached to biotic or abiotic surfaces embedded in a self-produced extracellular polymeric substance (EPS). The shift in bacteria from a free-swimming planktonic state to a biofilm-making sessile encompasses multiple stages: attachment to a surface, microcolony formation, biofilm maturation, and bacterial dispersion [7]. During the initial stage of biofilm formation, planktonic cells interact very briefly and transiently with a surface through Van der Waals forces by using flagella, pili, and fimbriae, depending on favorable environmental conditions. The gradual buildup in cellular cAMP levels and a gradual corresponding increase in type IV pili in the surface-associated cells make the attachment irreversible, and the cells start the production of an extracellular polymeric matrix composed of polysaccharides, proteins, and extracellular DNA (eDNA) that clump together, forming microcolonies [8].
EPS-embedded growing cells organize themselves into a three-dimensional structure, in which they interact, communicate, and cooperate through the complex communication system of QS. Lastly, intrinsic factors of the bacterial community, namely genetic changes, overpopulation, and intense competition for nutrients, as well as environmental perturbations like changes in nutrient concentrations and temperature, oxygen deficiency, and metabolite accumulation, lead to the dispersion of bacterial cells that return to a planktonic state [9].
The intrinsic properties of biofilms provide the microbial community with characteristics towards antibiotic resistance. It is interesting to consider Chen and Wen’s perspective, according to which bacterial biofilm is a particular kind of persistent bacterial infection [10]. Besides its structural role in biofilm, EPS represents a physical barrier for a biofilm since it hinders or reduces the penetration of antimicrobial compounds in the innermost layers. An example, ionic interactions between antimicrobial drugs and the EPS matrix may limit their penetration. The negative charge of polysaccharides effectively blocks the penetration of positively charged antibiotics such as aminoglycosides [11]. In addition, the EPS barrier leads to an incomplete penetration of antibiotics, and therefore, the cells in the biofilm layer encounter sub-inhibitory concentrations of the drug, becoming tolerant [12,13]. It has also been reported that extracellular DNA (eDNA), a key constituent of the EPS matrix, by acting as a metal cation chelator, activates the PhoPQ/PmrAB systems that control various genes involved in nutrient utilization, metal homeostasis, acid pH tolerance, and virulence in Pseudomonas aeruginosa [14]. The heterogeneity of the microenvironment within biofilm, characterized by a progressive gradient of pH, oxygen, and nutrients from the top to the bottom of the biofilm, also contributes to the acquisition of drug-resistant phenotypes. Andrè and coll. reported that pathogens benefit from hypoxia, which has been proven to activate various mechanisms of pathogen virulence [15]. An acidic microenvironment promotes faster bacterial evolution toward elevated antibiotic resistance, as evidenced by promoting the selection of stable genetic mutations and increasing the expressions of multiple biofilm- and virulence-related genes [16]. Further, in reduced levels of nutrients and oxygen, cells globally decrease their bacterial metabolic and growth rates, thus promoting the emergence of persister cells, a subpopulation of transiently slow-growing or growth-arrested cells that are tolerant to antibiotics and able to survive for extended periods in the absence of nutrients and lethal stresses [17,18]. The presence of persister cells can result in the recalcitrance and relapse of persistent bacterial infections, and it has been linked to an increase in the risk of the emergence of antibiotic resistance [19]. The increased resistance capacity of biofilms is also attributable to the upregulation of efflux pumps in response to antimicrobial exposure, thereby extruding multiple classes of antimicrobial compounds into the extracellular environment. A few efflux pump genes were found upregulated, mainly in the upper layer of biofilms. These include AcrAB-TolC in Escherichia coli, AdeFGH in Acinetobacter baumannii, PA1874-1877 and MexAB-OprM in P. aeruginosa, Resistance–Nodulation–cell Division (RND) in Burkholderia cenocepacia, and AcrD in Salmonella enterica [20,21,22,23]. Efflux pumps are also implicated in the secretion of QS molecules and thereby indirectly modulate gene expression remodeling underlying cell-to-cell adhesion, EPS and virulence factor production, and biofilm formation [24]. Lastly, within a biofilm, mobile genetic elements, including plasmids, transposons, and integrons, may facilitate the spread of resistance genes among the microbial community through horizontal gene transfer [25].

3. Anti-QS Approaches to Overcome Biofilm Resistance

Over the past two decades, several agents of diverse origins have been identified to suppress biofilm formation through the inhibition of the QS signaling pathway by exerting inhibitory effects at multiple levels. According to their target, these agents are generally grouped into three categories: inactivators of signaling molecules, inactivators of signaling receptors, and inhibitors of signaling cascade, through the mechanisms described below.

3.1. Inactivation of Signaling Molecules

Since inter- and intra-species bacterial communication is sensed, maintained, and powered by QS mechanisms and represents a crucial step for biofilm formation, an effective strategy to interrupt this communication is represented by the use of inhibitors capable of suppressing the synthesis of, inactivating, or degrading the AI signaling molecules. Studies on P. aeruginosa revealed that (z)-5-octylidenethiazolidine-2,4-dione (TZD-C8) significantly downregulates the expression of LuxI-type acyl-homoserine lactone synthases by interfering with both the Pseudomonas Quinolone Signal (PQS) and 3-oxo-C12-HSL signaling pathways [26]. In vitro studies also reported that subminimal growth-inhibitory concentrations of some macrolide antibiotics are able to reduce N-acyl homoserine lactone (AHL) synthesis in P. aeruginosa. Tateda and coll. revealed that azithromycin interferes with the synthesis of autoinducers by reducing the concentration of 3-oxo-C12-HSL and C4-HSL by 94 and 72%, respectively, and the expression of the transcriptional activator genes lasR and rhlR and of the two autoinducer synthase genes lasI and rhlI [27]. Subminimal growth-inhibitory concentrations of erythromycin were found to suppress the synthesis of homoserine lactone autoinducers as well as the production of hemagglutinins, proteases, and hemolysins [28].
Alternatively, AHL production could be blocked by targeting its two precursors, which are S-adenosylmethionine (SAM), an amino donor for the generation of the homoserine lactone ring moiety, and acyl carrier protein (ACP) for the acyl side chain of the AHL signal. Various analogues of SAM, such as S-adenosylhomocysteine, S-adenosylcysteine, and sinefungin, were demonstrated to be potent inhibitors of AHL synthesis [29]. A low concentration of azithromycin inhibits the expression of numerous genes belonging to the SAM synthesis pathway which, in turn, leads to a decrease in AHL production by LasI and RhlI [30]. What is more, triclosan reduces AHL synthesis by inhibiting the precursor of enoyl-ACP reductase [31].
A wide range of molecules have been progressively identified to inhibit AI-2 synthesis via the downregulation of the LuxS enzyme across a wide spectrum of bacteria. Next to S-homoribosyl-L-cysteine and L-homocysteine, sinefungin, a variety of halogenated furanones, and natural compounds like surfactin were recognized to downregulate the luxS, pfs, and speE genes, with a significant reduction in AI-2 production in several bacterial species, such as Streptococcus pneumoniae, Streptococcus suis, and Staphylococcus epidermidis [32,33,34,35].
Ultimately, the inactivation or degradation of signal molecules by AHL quenching represents another effective QQ strategy interfering with bacterial communication. In this context, three different classes of enzymes have been progressively discovered and classified according to their mechanism of action. These enzymes, their substrates, and relative effects on biofilm inhibition are reported in Table 1.
AHL lactonase hydrolyzes the ester bond of the homoserine lactone ring [71]. Most of the AHL-degrading enzymes are lactonases belonging to the metallo-lactamase superfamily [72]. Identified for the first time by Dong and coll. in Bacillus species isolated from soil samples, these enzymes have been progressively discovered in several other bacterial species, including A. tumefacens, Bacillus stearothermophilus, K. pneumoniae, Rhizobium sp., Mycobacterium avium, Muricauda olearia, and V. cholerae [6,36,38,43,45,46,51]. Most of the known AHL-lactonases have high affinity and stability at high temperatures and do not display any preference for the length of the carbon acyl side chain attached to the lactone ring [73].
Still identified in only a limited number of bacterial species, AHL acylases catalyze the degradation of the AHL amide bond, generating free fatty acids and a lattone ring [52,54,60,74,75]. It has been suggested that acylases degrade AHL with a long side chain than with a short side chain [76]. Lastly, several AHL oxidoreductases are known from various bacteria, reducing or oxidizing the acyl chain of the AHL, thus inhibiting the specific binding of the autoinducer to its receptor [65,66,69]. Only a few enzymes degrading or modifying AI-2 have been currently identified. Xavier and coll. identified in E. coli the kinase LsrK responsible for the phosphorylation of AI-2 as well as LsrG, potentially involved in the degradation of phospho-AI-2 [77,78]. In addition, since AI-2 in its dephosphorylated form is unable to bind to its receptor, and therefore the signaling cascade does not begin, it follows that many compounds exerting inhibition activity against LsrK have been progressively identified [79,80,81]. Furthermore, high concentrations of imidazole, a furan carbocyclic analogue of AI-2, decrease the antibiotic resistance of a clinical ampicillin-resistant E. coli strain by significantly downregulating the transcription of the lsrR gene that, in turn, reduces the function of AI-2 [82]. More recently, by using a metagenomic approach, several non-toxic biomolecules were identified. Particularly, it was demonstrated that the oxidoreductase QQ-2, simultaneously interfering with AI-2 and AHL, reduces 4-hydroxy-2,3-pentanedione-5-phosphate (P-DPD, a C5-phosphorylated derivative of the open AI-2 form) to 3,4,4-trihydroxy-2-pentanone-5-phosphate, namely, an inactive AI-2 derivative [70].

3.2. Inactivation of Signaling Receptors

In the last ten years, a few QS inhibitors competing with autoinducers for binding to their receptors have been discovered by computational docking studies as well as high-throughput screening (HTS) of small-molecule libraries, mainly in P. aeruginosa.
Flavonoids such as naringenin were demonstrated to directly bind the receptor LasR, thus preventing LasR/RhlR DNA binding, reducing biofilm formation, and repressing many other QS-related effects [83,84]. Similar QQ mechanisms were identified for many other natural compounds, including ortho-vanillin, embelin, and piperine, as well as for the antidiabetic drugs dipeptidase inhibitor-4 (DPI-4) sitagliptin and omarigliptin [85,86,87,88]. In addition, in vitro and molecular docking assays demonstrated that subinhibitory concentrations of several aminoglycoside drugs, including amikacin, gentamicin, kanamycin, neomycin B, paromomycin, and netilmicin, display strong binding properties with the LasR protein and, thus, exhibit antibiofilm action against P. aeruginosa [89]. More recently, Manson and coll., by screening a 25,000-compound library, discovered eight new robust and effective antagonists of LasR [90].
Numerous structural analogues of AHL were identified as antagonists of QS receptors to compete or interfere with the signal of autoinducers. In this context, several data demonstrated that the length of the acyl chain, as well as the introduction of an unsaturated bond close to the amide linkage of AHL, has a significant impact on its binding activity, so that AHL analogues with a long acyl side chain represent efficient inhibitors [91,92]. In addition, many natural compounds, including catechin, nakinadine B, flavonoids, and furanones, as well as many synthetic AHL analogues, were identified to exert a competitive inhibition on autoinducer receptors in several bacteria [84,93,94,95,96,97]. What is more, by using a multivalent scaffold approach, a chemical antagonist probe for Lsr-type AI-2 receptors in Salmonella typhimurium and an imaging agent for bacterial species utilizing Lsr-type AI-2 receptors were identified [98]. The antagonistic effect of several AI-2 synthetic analogues was recognized in Vibrio harveyi [99].

3.3. Inhibition of Signaling Cascade

The best-characterized mechanism of inhibition of the QS signaling cascades involves the use of targets blocking the downstream response regulator AgrA in S. aureus. Many compounds, such as Azan-7, bumetanide, savrin, and staquorsin, have been discovered to strongly reduce the AgrA-DNA complex formation at the P3 promoter region involved in the regulation of RNAIII transcription, thus preventing virulence gene upregulation and biofilm formation [100,101,102]. A few synthetic molecules have been discovered to act as antagonists of LuxR-type proteins, such as many autoinducer analogs. The TraR regulator response of Agrobacterium tumefaciens was demonstrated to be antagonized by analogs of the autoinducer 3-oxo-octanoyl-homoserine lactone [103]. Virstatin, a small molecule that prevents the expression of cholerae virulence factors, is able to repress the expression of AnoR that is a positive regulator of the LuxI-like synthase AnoI in Acinetobacter nosocomialis, leading to decreased synthesis of N-(3-hydroxy-dodecanoyl)-L-homoserine lactone (OH-dDHL). Low levels of this compound affect the signaling cascade and reduce biofilm formation and motility [104]. What is more, PAβN, an efflux pump inhibitor, reduces the extracellular accumulation of QS signaling molecules such as N-3-oxo-dodecanoyl-l-homoserine lactone and N-butyryl-l-homoserine lactone and significantly decreases the relative expression of the QS cascade (pqsA, pqsR, lasI, lasR, rhlI, and rhlR) and QS-regulated type II secretory genes lasB (elastase) and toxA (exotoxin A) in P. aeruginosa clinical isolates, thus inducing a reduction in bacterial virulence [105]. In E. coli and Salmonella, inhibitors of the transcription of the SdiA regulator protein, such as extracts of Melia dubia seeds and fructose-furoic acid, attenuate the expression of virulence factors by blocking the binding of AHLs to SdiA [106]. A phosphate ester derivative of chrysin exhibits higher anti-virulence activity by acting as a potent QS inhibitor of P. aeruginosa. It binds to the QS regulator LasR, thus abrogating its DNA-binding capability [107].
In Figure 1, the three different anti-QS approaches described above are schematized.

4. Quorum Quenching Compounds

Regarding their origin, QSIs can be distinguished into compounds of natural or synthetic provenance. Many QQ agents identified to date were isolated from various natural sources, such as plants, marine organisms, and living organisms.

4.1. Plant-Derived Anti-Biofilm Compounds

Plants have been known for centuries to be sources of a plethora of bioactive compounds with antioxidant, anticarcinogenic, antiallergenic, anti-inflammatory, antimutagenic, and antimicrobial activities, and it is no coincidence that phytochemicals represent the basis of traditional medicinal practices [108,109,110]. As part of the characterization of the antimicrobial properties of these compounds, referred to as phytochemicals, numerous studies have highlighted how the majority exert their action by significantly inhibiting the process of bacterial biofilm formation and the QS signaling system. Among these, we mention alkaloids, lectins, polyacetylenes, polyphenols, and terpenoids, of which a summary of the main compounds is shown in Table 2.
In recent years, special attention has been given to Essential Oils (EOs), volatile secondary metabolites of aromatic plants and spices that give them their characteristic and distinctive smell or taste, that exhibit recognized antimicrobial properties. In this frame, several studies have revealed how EOs can inhibit biofilm formation through different mechanisms. Calamintha nepeta (L.) Savi ssp. nepeta, Clinopodium nepeta L., Cinnamomum verum, Cinnamomum zeylanicum, Citrus bergamia, Citrus limon, Citrus reticulata, Eugenia caryophyllata, Foeniculum vulgare subsp. piperitum, Laurus nobilis L., Lippia alba, Myrtus communis L., Origanum vulgare spp., viridulum Ocimum basilicum, Prangos trifida, Petroselinum crispum, Salvia officinalis L., Salvia Rosmarinus, Satureja hortensis, Thymus daenensis, and Thymus vulgaris are just a few oils whose anti-biofilm properties have been demonstrated in both Gram-negative and Gram-positive bacteria as well as in multidrug-resistant clinical strains [137,138,139,140,141,142,143].
The biological activity of each EO has been mostly attributed to the hydrophobic nature of oils as well as to the action of their principal components, which are represented mainly by terpenes and terpenoids as primary components, and polyphenols [144]. It has been demonstrated that most of these compounds interact with the lipid bilayer and accumulate in the cell membrane, leading to a progressive loss of intracellular molecules such as nucleic acids and proteins [145,146]. Recent reports have also highlighted that EOs induce a profound altering of the global methylation levels of adenine and cytosine residues located in the genomes of both pathogenic and non-pathogenic bacterial strains, thus hypothesizing that they induce a massive epigenetic remodeling that, in turn, profoundly changes the expression of genes involved in cooperative behaviors, cell-to-cell adhesion and communication, and virulence and hinders the strategies adopted for the survival of the bacterial community [147].

4.2. Marine-Derived Anti-Biofilm Compounds

Also, the marine environment, with algae, invertebrates, sponges, and corals, represents a rich source of natural bioactive molecules; thus, in the last few years, their antibiofilm activity has been progressively identified. The inhibitory effects on biofilm formation exerted by extracts or derivatives from seaweeds and microalgae were recently reviewed by Behzadnia [148]. As an example, extracts from Westiellopsis prolifica inhibit biofilm formation and prevent cell adhesion against Gram-positive (S. aureus, Bacillus subtilis, and Streptococcus spp.) and Gram-negative bacteria (Shigella sp., Proteus sp., and P. aeruginosa), as well as fungi (Aspergills niger and C. albicans) [149]. Crude extracts from Arthrospira platensis (cyanobacteria) and Polysiphonia scopulorum (Rhodophyta) show strong antimicrobial and antibiofilm activity and are responsible for a significant downregulation of genes that play pivotal roles in surface attachment, biofilm formation, and the overall stability of the biofilm architecture in P. aeruginosa [150]. Similarly, Chlamydomonas extract inhibits the QS pathway, thus affecting P. aeruginosa pathogenicity and biofilm morphology and thickness [146]. Extracts from Cladostephus spongiosus (Phaeophyta) inhibit hyphal growth and biofilm formation via the downregulation of the expression of hyphal-specific genes and virulence factors in Candida sp. [151]. In addition, algal bioactive compounds have been progressively identified as anti-biofilm agents by mainly interfering with the expression of genes involved in QS signaling. Among these, we may mention algal polysaccharides, such as alginate, laminaran, and fucoidan, algal carotenoids, including zeaxanthin and lutein, algal lipids, of which α-linolenic acid, monogalactosylmonoacylglycerol, palmitoleic acid, spirulina, sulfoquinovosyldiacylglycerol, and sulfoquinovosylmonoacylglycerol are just some examples, as well as algal phlorotannins, that were found to reduce cell proliferation, motility, and adhesion and interfere with the synthesis of exopolysaccarides and virulence factors [152,153,154,155,156,157,158,159,160].
Furthermore, a few compounds from marine sponges were proven to have antibiofilm activity. These include the bromoageliferin from Agelas spp., the darwinolide from Dendrilla membranosa, the diterpene alkaloid(-)-ageloxime D from Agelas nakamurai, the tryptamine derivative bufotenine from Paramuricea clavate, and Synoxazolidinonones from Synoicum pulonaria [161]. QS inhibition associated with biofilm disruption was observed in Vibrio spp. treated with a marine actinomyces extract [162].

4.3. Microbial-Derived Anti-Biofilm Compounds

Next to the previously described enzymes involved in the inactivation or degradation of QS autoinducers, namely lactonases, acylases, and oxidoreductases, and to the well-characterized antibiotics obtained from microorganisms, such as benzylpenicillin, cephalosporins, and gentamicin, many secondary microbial metabolites have been progressively identified to exert QQ activity [163]. Walkmycin C, a histidine kinase (HK) inhibitor produced by Streptomyces sp., displays anti-biofilm activity against Streptococcus mutans [164]. In addition, alnumycin D, granaticin B, kalafungin, medermycin, collismycin C, napyradiomycin, hygrocin C, 8-O-metyltetrangomycin, panglimycin D, and streptorubin, isolated from Streptomyces sp. exhibit promising antibiofilm activities against S. aureus and methicillin-resistant S. aureus (MRSA) [165,166,167,168,169].
Several microbial-produced diffusible-signal factors (DSF), which are cis-2-unsaturated fatty acids identified in a range of bacterial pathogens, including Xanthomonas, Enterobacter, Thiobacillus, Leptospirillum, Stenotrophomonas, Burkholderia, Achromobacter, Yersinia, Methylobacillus, Pantoea, and Cronobacter, are able to disperse biofilm formation in a range of Gram-negative and Gram-positive bacteria and modulate virulence gene expression, reverse persistence, increase microbial metabolic activity, and significantly enhance the antibacterial activity of conventional antimicrobial agents [170,171,172,173]. Carolacton, a macrolide keto-carboxylic acid produced by the myxobacterium Sorangium cellulosum, by interfering with PknB-mediated signaling in growing cells, significantly inhibits S. mutans biofilm formation [174]. Similarly, other secondary metabolite compounds, including the skyllamycin family produced by Streptomyces sp., phenazines from both Gram-positive and Gram-negative bacteria, and promysalin isolated by Pseudomonas putida, were demonstrated to exert an antibiofilm effect [175].
Biosurfactants (BSs) are amphipathic bioactive compounds that are produced mostly by microorganisms on their cell surface or secreted extracellularly, and they are classified, according to the chemical structure of the hydrophilic head group, in glycolipids, fatty acids, lipopeptides, and polymers. They are powerful anti-biofilm agents that can act as biocides as well as biodispersants up to 95.9% [176,177].
Rhamnolipids, glycolipids, surfactins, and lipopepdide biosurfactants exhibit antimicrobial, anti-fungal, and anti-biofilm activity through different mechanisms of action. In vitro and ex vivo evidence reports that a biosurfactant from the entomopathogenic fungus Beauveria bassiana alters cell permeability and disrupts the cell membrane, effectively eradicating the biofilm of Microsporum canis [178]. Some BSs, including surfactin, rhamnolipid, and lipopeptide pontifactin, alter the hydrophobicity of the cell surface, interfere with cell adhesion to other microorganisms and/or surfaces, and affect the integrity of cellular membranes, thereby dispersing the biofilm and solubilizing the component of the matrix [179,180,181]. Lastly, BSs may indirectly inhibit the biofilm by interfering with QS signals. In this context, Yan and coll. demonstrated that BSs isolated from lactic acid bacteria, including Pediococcus acidilactici and Lactobacillus plantarum, affect the expression of genes involved in biofilm formation (cidA, icaA, dltB, agrA, sortaseA, and sarA) and interfere with the release of AI-2 signaling factor in S. aureus [182].

4.4. Antimicrobial Peptides

Antimicrobial peptides (AMPs) are a class of agents naturally occurring in animals, plants, and microbes, exhibiting antimicrobial, anti-attachment, and antibiofilm properties against both Gram-positive and Gram-negative multidrug-resistant pathogens and fungi as well [183,184]. Most AMPs are cationic, which allows their interaction with the negatively charged bacterial membranes, which permealize, thus leading to cell death [185,186].
Evidence has reported that AMPs affect biofilm formation or degradation at different stages and through the following mechanisms of action:
(i)
Inhibition of bacterial attachments to surfaces: AMPs can interfere in the early stages of biofilm formation to prevent the initial adhesion of bacteria to surfaces. Arslan and coll. reported that lactoferrin suppresses the initial attachment of Streptococcus gordonii coaggregates [187]. By downregulating genes encoding ABC transporters involved in cell-to-surface and cell-to-cell interactions, the peptide Nal-P-113 can inhibit Porphyromonas gingivalis biofilm formation [188].
(ii)
Disruption of the membrane potential of biofilm-embedded cells: Different kinds of bacteriocins, including nisin A and, to a lesser extent, lacticin Q, show antibiofilm activity on clinical isolates of methicillin-resistant S. aureus (MRSA), altering the membrane potential [189].
(iii)
Permeabilization of cell membrane: Esculentins Esc (1–21) and (CSA)-13 permeabilize the cytoplasmic membrane of P. aeruginosa in biofilms [190,191].
(iv)
Interruption of QS signaling and gene regulation: In P. aeruginosa, cathelicidic LL-37 and indolicidin prevent biofilm formation via the downregulation of the transcription of the Las and Rhl systems [192]. Additionally, peptide 1037 directly inhibits biofilms by reducing swimming and swarming motilities, stimulating twitching motility, and suppressing the expression of a variety of genes involved in biofilm formation [193]. The human β-defensin 3 (hBD-3) reduces the synthesis of polysaccharide intercellular adhesin (PIA) by downregulating the expression of the icaA, icaD, and icaR genes in S. epidermidis [194].
(v)
Degradation of the biofilm matrix: Peptide PI decrease S. mutans biofilm biomass by degrading the exopolysaccharide matrix [195]. In addition, an AMP complex produced by the maggots of blowfly Calliphora vicina effectively counteracts the formation of E. coli, S. aureus, and A. baumannii biofilms by destroying the biofilm matrix [196]. Similarly, the human antimicrobial peptide hepcidin 20 reduces the mass of the extracellular matrix and alters the architecture of the biofilm in S. epidermidis [197].
Lastly, bacteriocins are a group of antimicrobial peptides produced by bacteria that have the potential to disrupt biofilm either by themselves or in combination with other drugs [198]. Evidence has reported that sonorensin, a bacteriocin isolated from Bacillus sonorensis, inhibits the biofilm growth of S. aureus by increasing membrane permeability [199]. The bacteriocin LFX01A from Lactiplantibacillus plantarum has antibiofilm activity against Shigella flexneri [200]. Bacteriocins produced by Pediococcus acidilactici exhibit antibiofilm activity through a decrease in extracellular polymeric substances (EPSs) and a reduction in cell adhesion [201]. Bacteriocin from Lactobacillus plantarum was effective in counteracting biofilm formation on catheters induced by bacteria such as P. aeruginosa and S. aureus [202].

4.5. Synthetic Compounds

A plethora of artificially designed inhibitory molecules are being synthesized to target QS and biofilm processes in bacteria. These molecules, deeply detailed by Vashistha and coll., primarily include analogues of autoinducers of QS that competitively bind the AHL receptors, thus inhibiting QS and biofilm formation [203]. Next to these analogues, over the past 15 years, several studies have highlighted the promising potential of synthetic nanostructured materials for counteracting bacterial growth. Nanoparticles (NPs) such as gold (Au), silver (Ag), copper (Cu), palladium (Pd), platinum (Pt), or zinc oxide (ZnO) have been proposed as novel potential means of fighting bacteria thanks to their ability to induce death or limit the growth of microorganisms depending on their particular physicochemical parameters, namely the shape, size, composition, and surface functionalization [204,205,206,207]. Their high surface area-to-volume ratio, high reactivity, and stability, combined with their unique physicochemical properties, allow them to interact with microbial membranes and exhibit antimicrobial effects at low concentrations by inducing oxidative stress and disrupting membrane integrity and bacterial biomolecules [206,208]. These same parameters were also found to be fundamental in conferring antibiofilm properties to NPs. More particularly, the interaction between NPs and bacteria embedded in a biofilm is modulated by the physicochemical properties of the NPs, such as their size, charge, shape, and hydrophobicity and the EPS matrix [209]. The smaller the size of the NPs, the greater their ability to penetrate the water-filled channels of the biofilm and exert antibiofilm and antibacterial effects [210,211,212]. Likewise, shapes that allow for a high surface area-to-volume ratio give NPs greater antibacterial and anti-biofilm efficacy, with rod-like NPs exhibiting more effectiveness than spherical particles in their inhibitory action [211]. In addition, the hydrophobicity of the NP surface determines its efficacy in penetrating into biofilm. Li and coll. demonstrated that NPs coated with DSPE-PEG2000 increase their hydrophilicity, improve their penetration through mucus, and effectively eradicate Helicobacter pylori biofilm [213]. It has also been reported that cationic particles penetrate biofilms completely, while neutral or anionic ones are unable to do so, since the EPS substances as well as the bacterial cell wall have an overall negative charge [214]. On the other hand, the viscosity and compaction of the EPS matrix, cell density, and the characteristics of the water spaces (pores) within the EPS matrix influence the speed and penetration of NPs within a biofilm [215]. A summary of the different types of nanocomposites, their characteristics, and their antibiofilm activities has been recently published [216]. NPs can exert their antibiofilm activity via a multitude of mechanisms [217,218]. The antibacterial and antibiofilm properties of silver nanoparticles (AgNPs) have been investigated in several studies. Evidence demonstrates that AgNPs fuse with the cell membrane, altering its charge and affecting cell permeability, thus ultimately leading to cell lysis [219]. These particles are also able to downregulate the production of exopolysaccharide and alginate as well as interfere with the QS system [220]. The antibiofilm effect brought out by gold nanoparticles (AuNPs), which is enhanced by the intrinsic photothermal properties of this metal, is primarily amenable to the reduction in exopolysaccharide production, the inhibition of swimming and swarming motility, and the disruption of the cell membrane followed by a metabolic deterioration [218,221,222]. Zinc nanoparticles (ZnNPs) induce cell depolarization, increase permeability, increase oxidative stress, and alter the expression of QS genes [223]. Similarly, Superparamagnetic Iron Oxide Nanoparticles (SPIONs) have been shown to reduce biofilm formation in S. mutans with different effectiveness depending on their surface charge [224]. Being magnetic, the presence of an external magnetic field induces these NPs to penetrate the full depth of the biofilm, thus killing bacterial cells deep inside biofilms due to vibration damage, local hyperthermia, and ROS generation [225,226].
If nanomaterials themselves act by inhibiting biofilms quite effectively through the above-mentioned mechanisms, their action is even more enhanced when they are conjugated together to form bimetallic nanoparticles (BMNPs) or coated with QS inhibitors. As an example, studies carried out on B. subtilis, Proteus mirabilis, and S. mutans revealed that the antibiofilm effect of mixed ZnO:MgO NPs and Zn:CuO NPs is more prominent compared to that shown by the individual NPs [227,228]. Ho and coll. modified polyacrylamide magnetic beads with the funarone-like analogue Dihydropyrrolidone (DHP), a molecule with proven QQ activity, thus revealing the effectiveness of the substrate in reducing the growth of S. aureus in vitro and in subcutaneous infection models [229]. Similarly, nanocarriers of many other QQ compounds, including flavonoids, AMPs, and autoinducer-degrading enzymes, gave encouraging results in promoting the dispersion of biofilms. NPCs loaded with a combination of tt-farnesol and compound 1771, with or without myricetin, fully prevent S. mutans and C. albicans dual-species biofilm formation by impeding biomass accumulation, bacterial growth, and exopolysaccharide matrix deposition [230]. AuNPs coated with the lactonase AiiA appeared to be effective in degrading K-hexanoyl-L-homoserine lactone, suppressing EPS production, and biofilm formation in multidrug-resistant Proteus species [231].
Lastly, lipid-coated hybrid nanoparticles (LCHNPs) were investigated for their capability of delivering antibiotics into biofilms. In vitro experiments demonstrated that LCHNPs loaded with vancomycin eradicate up to 99.99% of S. aureus biofilm, an effect that is not observed following treatment with vancomycin alone [232]. NPs encapsulated with ciprofloxacin and polymyxin B show complete inhibition against E. coli and S. aureus growth and P. aeruginosa biofilm [233,234]. Tobramycin and an alkylquinolone QS inhibitor encapsulated in squalenyl hydrogen sulfate nanoparticles (SqNPs) allow for the complete eradication of biofilm at a 16-fold lower concentration than tobramycin alone [235]. Hybrid nanospheres containing gentamicin and QQ acylase exhibit a significant inhibitory effect on the formation of P. aeruginosa biofilm [236].

5. QQ Applications

In concert with the progress of knowledge regarding the molecular basis of the QQ process, innovative applications of QS inhibitors are providing encouraging results in mitigating the negative impact that microorganisms have in different fields such as agriculture, aquaculture, waste treatment, and human health.
Bacterial soft rot is a most pervasive and economically pernicious disease that affects a wide range of agricultural and horticultural crops and is caused by multiple genera of Gram-positive and Gram-negative phytopathogens [237]. In exploring pesticide-substitutive and environmentally friendly strategies that may finally allow for effective disease management, a QQ-based approach, mainly based on the identification of AHL-degrading strains, is gaining considerable attention. The Pseudomonas segetis strain P6 isolated from the halophyte plant Salicornia europaea displays biocontrol ability against the phytopathogens Dickeya solani, Pectobacterium atrosepticum, P. carotovorum subsp. carotovorum, and Pseudomonas syringae infecting tomato and carrot plants by degrading AHL [238]. Bacillus thuringiensis strains possess the AHL-lactonase AiiA, which suppresses the QS-dependent virulence of the plant pathogen E. carotovora, thereby reducing its pathogenicity and the incidence of the symptom development of potato soft rot [239]. Zang and coll. identified a highly efficient AHL-degrading Pseudomonas nitroreducens strain W-7 that is capable of degrading a wide range of AHLs, including N-(3-oxohexanoyl)-l-homoserine lactone (OHHL), N-(3-oxooctanoyl)-l-homoserine lactone (OOHL), and N-hexanoyl-l-homoserine lactone (HHL). It follows that this strain is a useful biocontrol agent against various bacterial phytopathogens [240]. Bacterial strains capable of degrading the Diffusible Signal Factor (DSF) involved in the regulation of pathogenic virulence have also been identified. Among these, we mention Acinetobacter lactucae strain QL-1, Burkholderia anthina strain HN-8, and Burkholderia sp. F25, characterized for their degradation ability and potential biocontrol of black rot disease caused by Xanthomonas campestris [241,242,243]. In addition to this biocontrol approach, transgenic plants modified with QQ lactonase have also been developed to prevent bacterial infection. Dong and coll. first demonstrated that tobacco and potato plants expressing aiiA lactonase from Bacillus sp. quench pathogen QS signaling and show significantly enhanced resistance to E. carotovora infection [71]. Nicotiana tabacum lines expressing the lactonase AttM efficiently quench P. carotovorum communication in vivo [244].
Stress conditions (due to overcrowding) and water quality in aquaculture make fish particularly susceptible to bacterial infections, mainly from Vibrio, Pseudomonas, Edwardsiella, Flavobacterium, and Aeromonas. In the last few years, non-conventional methods for controlling bacterial diseases and improving the general health conditions of cultured species have been adopted to attenuate the virulence of fish pathogens. Lactonases purified from Bacillus sp. can attenuate A. hydrophila infections in zebrafish [245]. The AiiA lactonase isolated from Bacillus licheniformis DAHB1 inhibits in vitro biofilm formation in Vibrio and reduces its colonization and the mortality of shrimps in aquaculture [246]. Finally, Bacillus sp. QSI-1, B. licheniformis T-1, B. thuringensis QQ1, B. cereus QQ2, and Bacillus velezensis D-18 significantly reduce the pathogenicity of A. hydrophila and biofilm formation in Vibrio sp. by downregulating the production of the respective virulence factors, thereby making them safe and effective for protecting hosts against pathogenic bacterial infections in aquaculture [247,248,249,250].
Membrane bioreactors (MBRs) represent the technology of choice to treat both industrial and municipal wastewaters. However, the accumulation of microbial cells and extracellular polymeric substances, namely, biofouling, represents the critical point that negatively impacts the filtration performance of the membranes. Therefore, the current non-chemical strategies for reducing membrane biofouling consist of the immobilization of QQ bacteria, QQ enzymes, or QQ natural compounds such as curcumin, flavonoids, and funarones on different surfaces to significantly inhibit biofilm formation [251,252,253].
Biofilm formation by pathogenic bacteria on medical devices and implants generally evolves into persistent and often recurrent infections that are often refractory to conventional antibiotic treatments and, therefore, challenging to treat. As elaborated upon in the previous paragraph, nanomaterials per se or in combination with QS inhibitors represent an innovative and effective solution to prevent, delay, and eradicate biofilm formation on surfaces or for enhancing drug targeting.

6. Conclusions

Currently, novel effective therapeutical approaches are being feverishly searched for to deal with the emergence and spread of antibiotic-resistant pathogens. In this context, a better characterization of the molecular mechanisms underlying QS and its inhibition, as well as the discovery of new compounds with quenching activity, turns out to be fundamental for the development of innovative and antibiotic-free multifaceted biocontainment approaches in different fields such as the clinical, food, and environmental agriculture fields.

Author Contributions

Writing—Original Draft Preparation, D.B. and P.D.; Writing—Review and Editing, P.D., E.D.R., G.S., A.S., B.C., G.P. and D.B.; Funding Acquisition, G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Next Generation EU in the context of the National Recovery and Resilience Plan, Investment PE8—Project Age-It: “Ageing Well in an Ageing Society”. This resource was co-financed by the Next Generation EU [DM 1557 11.10.2022]. This work has been made possible by collaboration with Gruppo Baffa (Sadel Spa, Sadel San Teodoro srl, Sadel CSsrl, Casa di Cura Madonna dello Scoglio, AGI srl, Casa di Cura Villa del Rosario srl, Savelli Hospital srl) and Casa di Cura Villa Ermelinda in the frame of the agreements TECNOLOGIE AVANZATE PER IL MIGLIORAMENTO DEI SERVIZI ALLA PERSONA and TECNOLOGIE AVANZATE PER L’INDAGINE DELLE RELAZIONI TRA UOMO ED AMBIENTI DI VITA.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, L.; Luo, Y. Bacterial Quorum-Sensing Systems and Their Role in Intestinal Bacteria-Host Crosstalk. Front. Microbiol. 2021, 12, 611413. [Google Scholar] [CrossRef] [PubMed]
  3. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef] [PubMed]
  4. Sharma, D.; Misba, L.; Khan, A.U. Antibiotics versus biofilm: An emerging battleground in microbial communities. Antimicrob. Resist. Infect. Control 2019, 8, 76. [Google Scholar] [CrossRef]
  5. Jamal, M.; Ahmad, W.; Andleeb, S.; Jalil, F.; Imran, M.; Nawaz, M.A.; Hussain, T.; Ali, M.; Rafiq, M.; Kamil, M.A. Bacterial biofilm and associated infections. J. Chin. Med. Assoc. 2018, 81, 7–11. [Google Scholar] [CrossRef]
  6. Zhu, X.; Chen, W.J.; Bhatt, K.; Zhou, Z.; Huang, Y.; Zhang, L.H.; Chen, S.; Wang, J. Innovative microbial disease biocontrol strategies mediated by quorum quenching and their multifaceted applications: A review. Front. Plant Sci. 2023, 13, 1063393. [Google Scholar] [CrossRef]
  7. Rather, M.A.; Gupta, K.; Mandal, M. Microbial biofilm: Formation, architecture, antibiotic resistance, and control strategies. Braz. J. Microbiol. 2021, 52, 1701–1718. [Google Scholar] [CrossRef] [PubMed]
  8. Armbruster, C.R.; Parsek, M.R. New insight into the early stages of biofilm formation. Proc. Natl. Acad. Sci. USA 2018, 115, 4317–4319. [Google Scholar] [CrossRef]
  9. McDougald, D.; Rice, S.A.; Barraud, N.; Steinberg, P.D.; Kjelleberg, S. Should we stay or should we go: Mechanisms and ecological consequences for biofilm dispersal. Nat. Rev. Microbiol. 2011, 10, 39–50. [Google Scholar] [CrossRef]
  10. Chen, L.; Wen, Y.M. The role of bacterial biofilm in persistent infections and control strategies. Int. J. Oral Sci. 2011, 3, 66–73. [Google Scholar] [CrossRef]
  11. Tseng, B.S.; Zhang, W.; Harrison, J.J.; Quach, T.P.; Song, J.L.; Penterman, J.; Singh, P.K.; Chopp, D.L.; Packman, A.I.; Parsek, M.R. The extracellular matrix protects Pseudomonas aeruginosa biofilms by limiting the penetration of tobramycin. Environ. Microbiol. 2013, 15, 2865–2878. [Google Scholar] [CrossRef] [PubMed]
  12. Sulaiman, J.E.; Lam, H. Evolution of Bacterial Tolerance Under Antibiotic Treatment and Its Implications on the Development of Resistance. Front. Microbiol. 2021, 12, 617412. [Google Scholar] [CrossRef] [PubMed]
  13. Kosztołowicz, T.; Metzler, R. Diffusion of antibiotics through a biofilm in the presence of diffusion and absorption barriers. Phys. Rev. E 2020, 102, 032408. [Google Scholar] [CrossRef] [PubMed]
  14. Lewenza, S.; Johnson, L.; Charron-Mazenod, L.; Hong, M.; Mulcahy-O’Grady, H. Extracellular DNA controls expression of Pseudomonas aeruginosa genes involved in nutrient utilization, metal homeostasis, acid pH tolerance and virulence. J. Med. Microbiol. 2020, 69, 895–905. [Google Scholar] [CrossRef]
  15. André, A.C.; Laborde, M.; Marteyn, B.S. The battle for oxygen during bacterial and fungal infections. Trends Microbiol. 2022, 30, 643–653. [Google Scholar] [CrossRef]
  16. Lin, Q.; Pilewski, J.M.; Di, Y.P. Acidic Microenvironment Determines Antibiotic Susceptibility and Biofilm Formation of Pseudomonas aeruginosa. Front. Microbiol. 2021, 12, 747834. [Google Scholar] [CrossRef]
  17. Kouhsari, E.; Kaviar, V.H.; Asadi, A.; Ahmadi, A.; Sholeh, M.; Mirbalouchzehi, A.; Yaghoubi, S.; Abdi, M. Bacterial Persister Cells: Mechanisms of Formation, Control, and Eradication. Infect. Disord. Drug Targets 2023, 23, 17–28. [Google Scholar] [CrossRef]
  18. Fisher, R.A.; Gollan, B.; Helaine, S. Persistent bacterial infections and persister cells. Nat. Rev. Microbiol. 2017, 15, 453–464. [Google Scholar] [CrossRef]
  19. Zhou, Y.; Liao, H.; Pei, L.; Pu, Y. Combatting persister cells: The daunting task in post-antibiotics era. Cell Insight 2023, 2, 100104. [Google Scholar] [CrossRef]
  20. Alav, I.; Sutton, J.M.; Rahman, K.M. Role of bacterial efflux pumps in biofilm formation. J. Antimicrob. Chemother. 2018, 73, 2003–2020. [Google Scholar] [CrossRef]
  21. He, X.; Lu, F.; Yuan, F.; Jiang, D.; Zhao, P.; Zhu, J.; Cheng, H.; Cao, J.; Lu, G. Biofilm Formation Caused by Clinical Acinetobacter baumannii Isolates Is Associated with Overexpression of the AdeFGH Efflux Pump. Antimicrob. Agents Chemother. 2015, 59, 4817–4825. [Google Scholar] [CrossRef] [PubMed]
  22. Buroni, S.; Matthijs, N.; Spadaro, F.; Van Acker, H.; Scoffone, V.C.; Pasca, M.R.; Riccardi, G.; Coenye, T. Differential roles of RND efflux pumps in antimicrobial drug resistance of sessile and planktonic Burkholderia cenocepacia cells. Antimicrob. Agents Chemother. 2014, 58, 7424–7429. [Google Scholar] [CrossRef]
  23. Liao, J.; Schurr, M.J.; Sauer, K. The MerR-like regulator BrlR confers biofilm tolerance by activating multidrug efflux pumps in Pseudomonas aeruginosa biofilms. J. Bacteriol. 2013, 195, 3352–3363. [Google Scholar] [CrossRef] [PubMed]
  24. Hajiagha, M.N.; Kafil, H.S. Efflux pumps and microbial biofilm formation. Infect. Genet. Evol. 2023, 112, 105459. [Google Scholar] [CrossRef]
  25. Michaelis, C.; Grohmann, E. Horizontal Gene Transfer of Antibiotic Resistance Genes in Biofilms. Antibiotics 2023, 12, 328. [Google Scholar] [CrossRef] [PubMed]
  26. Lidor, O.; Al-Quntar, A.; Pesci, E.C.; Steinberg, D. Mechanistic analysis of a synthetic inhibitor of the Pseudomonas aeruginosa LasI quorum-sensing signal synthase. Sci. Rep. 2015, 5, 16569. [Google Scholar] [CrossRef]
  27. Tateda, K.; Comte, R.; Pechere, J.C.; Köhler, T.; Yamaguchi, K.; Van Delden, C. Azithromycin inhibits quorum sensing in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2001, 45, 1930–1933. [Google Scholar] [CrossRef] [PubMed]
  28. Sofer, D.; Gilboa-Garber, N.; Belz, A.; Garber, N.C. ‘Subinhibitory’ erythromycin represses production of Pseudomonas aeruginosa lectins, autoinducer and virulence factors. Chemotherapy 1999, 45, 335–341. [Google Scholar] [CrossRef]
  29. Parsek, M.R.; Val, D.L.; Hanzelka, B.L.; Cronan, J.E., Jr.; Greenberg, E.P. Acyl homoserine-lactone quorum-sensing signal generation. Proc. Natl. Acad. Sci. USA 1999, 96, 4360–4365. [Google Scholar] [CrossRef]
  30. Kai, T.; Tateda, K.; Kimura, S.; Ishii, Y.; Ito, H.; Yoshida, H.; Kimura, T.; Yamaguchi, K. A low concentration of azithromycin inhibits the mRNA expression of N-acyl homoserine lactone synthesis enzymes, upstream of lasI or rhlI, in Pseudomonas aeruginosa. Pulm. Pharmacol. Ther. 2009, 22, 483–486. [Google Scholar] [CrossRef]
  31. Hoang, T.T.; Schweizer, H.P. Characterization of Pseudomonas aeruginosa enoyl-acyl carrier protein reductase (FabI): A target for the antimicrobial triclosan and its role in acylated homoserine lactone synthesis. J. Bacteriol. 1999, 181, 5489–5497. [Google Scholar] [CrossRef]
  32. Yadav, M.K.; Park, S.W.; Chae, S.W.; Song, J.J. Sinefungin, a natural nucleoside analogue of S-adenosylmethionine, inhibits Streptococcus pneumoniae biofilm growth. BioMed Res. Int. 2014, 2014, 156987. [Google Scholar] [CrossRef]
  33. Han, X.; Lu, C. Biological activity and identification of a peptide inhibitor of LuxS from Streptococcus suis serotype 2. FEMS Microbiol. Lett. 2009, 294, 16–23. [Google Scholar] [CrossRef] [PubMed]
  34. Zang, T.; Lee, B.W.; Cannon, L.M.; Ritter, K.A.; Dai, S.; Ren, D.; Wood, T.K.; Zhou, Z.S. A naturally occurring brominated furanone covalently modifies and inactivates LuxS. Bioorg. Med. Chem. Lett. 2009, 19, 6200–6204. [Google Scholar] [CrossRef] [PubMed]
  35. Benneche, T.; Hussain, Z.; Aamdal Scheieb, A.; Lönn-Stensrudb, J. Synthesis of 5-(bromomethylene) furan-2(5H)-ones and 3-(bromomethylene)isobenzofuran-1(3H)-ones as inhibitors of microbial quorum sensing. New J. Chem. 2008, 32, 1567–1572. [Google Scholar] [CrossRef]
  36. Shevate, S.N.; Shinde, S.S.; Bankar, A.V.; Patil, N.P. Identification of Quorum Quenching N-Acyl Homoserine Lactonases from Priestia aryabhattai J1D and Bacillus cereus G Isolated from the Rhizosphere. Curr. Microbiol. 2023, 80, 86. [Google Scholar] [CrossRef] [PubMed]
  37. Rajesh, P.S.; Rai, V.R. Inhibition of QS-regulated virulence factors in Pseudomonas aeruginosa PAO1 and Pectobacterium carotovorum by AHL-lactonase of endophytic bacterium Bacillus cereus VT96. Biocatal. Agric. Biotechnol. 2016, 7, 154–163. [Google Scholar] [CrossRef]
  38. Carlier, A.; Uroz, S.; Smadja, B.; Fray, R.; Latour, X.; Dessaux, Y.; Faure, D. The Ti plasmid of Agrobacterium tumefaciens harbors an attM-paralogous gene, aiiB, also encoding N-Acyl homoserine lactonase activity. Appl. Environ. Microbiol. 2003, 69, 4989–4993. [Google Scholar] [CrossRef]
  39. Dong, W.; Cai, Y.; Xu, Z.; Fu, B.; Chen, Q.; Cui, Y.; Ruan, Z.; Liang, Y.; Peng, N.; Zhao, S. Heterologous expression of AHL lactonase AiiK by Lactobacillus casei MCJΔ1 with great quorum quenching ability against Aeromonas hydrophila AH-1 and AH-4. Microb. Cell Fact. 2020, 19, 191. [Google Scholar] [CrossRef]
  40. Dong, W.; Zhu, J.; Guo, X.; Kong, D.; Zhang, Q.; Zhou, Y.; Liu, X.; Zhao, S.; Ruan, Z. Characterization of AiiK, an AHL lactonase, from Kurthia huakui LAM0618T and its application in quorum quenching on Pseudomonas aeruginosa PAO1. Sci. Rep. 2018, 8, 6013. [Google Scholar] [CrossRef]
  41. Fan, X.; Liang, M.; Wang, L.; Chen, R.; Li, H.; Liu, X. Aii810, a Novel Cold-Adapted N-Acylhomoserine Lactonase Discovered in a Metagenome, Can Strongly Attenuate Pseudomonas aeruginosa Virulence Factors and Biofilm Formation. Front. Microbiol. 2017, 8, 1950. [Google Scholar] [CrossRef]
  42. Schipper, C.; Hornung, C.; Bijtenhoorn, P.; Quitschau, M.; Grond, S.; Streit, W.R. Metagenome-derived clones encoding two novel lactonase family proteins involved in biofilm inhibition in Pseudomonas aeruginosa. Appl. Environ. Microbiol. 2009, 75, 224–233. [Google Scholar] [CrossRef]
  43. Krysciak, D.; Schmeisser, C.; Preuss, S.; Riethausen, J.; Quitschau, M.; Grond, S.; Streit, W.R. Involvement of multiple loci in quorum quenching of autoinducer I molecules in the nitrogen-fixing symbiont Rhizobium (Sinorhizobium) sp. strain NGR234. Appl. Environ. Microbiol. 2011, 77, 5089–5099. [Google Scholar] [CrossRef]
  44. Chow, J.Y.; Yang, Y.; Tay, S.B.; Chua, K.L.; Yew, W.S. Disruption of biofilm formation by the human pathogen Acinetobacter baumannii using engineered quorum-quenching lactonases. Antimicrob. Agents Chemother. 2014, 58, 1802–1805. [Google Scholar] [CrossRef]
  45. Chow, J.Y.; Wu, L.; Yew, W.S. Directed evolution of a quorum-quenching lactonase from Mycobacterium avium subsp. paratuberculosis K-10 in the amidohydrolase superfamily. Biochemistry 2009, 48, 4344–4353. [Google Scholar] [CrossRef] [PubMed]
  46. Tang, K.; Su, Y.; Brackman, G.; Cui, F.; Zhang, Y.; Shi, X.; Coenye, T.; Zhang, X.H. MomL, a novel marine-derived N-acyl homoserine lactonase from Muricauda olearia. Appl. Environ. Microbiol. 2015, 81, 774–782. [Google Scholar] [CrossRef] [PubMed]
  47. Bzdrenga, J.; Daudé, D.; Rémy, B.; Jacquet, P.; Plener, L.; Elias, M.; Chabrière, E. Biotechnological applications of quorum quenching enzymes. Chem. Biol. Interact. 2017, 267, 104–115. [Google Scholar] [CrossRef]
  48. Chun, C.K.; Ozer, E.A.; Welsh, M.J.; Zabner, J.; Greenberg, E.P. Inactivation of a Pseudomonas aeruginosa quorum-sensing signal by human airway epithelia. Proc. Natl. Acad. Sci. USA 2004, 101, 3587–3590. [Google Scholar] [CrossRef] [PubMed]
  49. Curcic, J.; Dinic, M.; Novovic, K.; Vasiljevic, Z.; Kojic, M.; Jovcic, B.; Malesevic, M. A novel thermostable YtnP lactonase from Stenotrophomonas maltophilia inhibits Pseudomonas aeruginosa virulence in vitro and in vivo. Int. J. Biol. Macromol. 2024, 264, 130421. [Google Scholar] [CrossRef]
  50. Chen, C.N.; Chen, C.J.; Liao, C.T.; Lee, C.Y. A probable aculeacin A acylase from the Ralstonia solanacearum GMI1000 is N-acyl-homoserine lactone acylase with quorum-quenching activity. BMC Microbiol. 2009, 9, 89. [Google Scholar] [CrossRef]
  51. Park, S.Y.; Lee, S.J.; Oh, T.K.; Oh, J.W.; Koo, B.T.; Yum, D.Y.; Lee, J.K. AhlD, an N-acylhomoserine lactonase in Arthrobacter sp., and predicted homologues in other bacteria. Microbiology 2003, 149, 1541–1550. [Google Scholar] [CrossRef] [PubMed]
  52. Park, S.Y.; Kang, H.O.; Jang, H.S.; Lee, J.K.; Koo, B.T.; Yum, D.Y. Identification of extracellular N-acylhomoserine lactone acylase from a Streptomyces sp. and its application to quorum quenching. Appl. Environ. Microbiol. 2005, 71, 2632–2641. [Google Scholar] [CrossRef] [PubMed]
  53. Terwagne, M.; Mirabella, A.; Lemaire, J.; Deschamps, C.; De Bolle, X.; Letesson, J.J. Quorum sensing and self-quorum quenching in the intracellular pathogen Brucellamelitensis. PLoS ONE 2013, 8, e82514. [Google Scholar] [CrossRef] [PubMed]
  54. Lin, Y.H.; Xu, J.L.; Hu, J.; Wang, L.H.; Ong, S.L.; Leadbetter, J.R.; Zhang, L.H. Acyl-homoserine lactone acylase from Ralstonia strain XJ12B represents a novel and potent class of quorum-quenching enzymes. Mol. Microbiol. 2003, 47, 849–860. [Google Scholar] [CrossRef] [PubMed]
  55. Czajkowski, R.; Krzyżanowska, D.; Karczewska, J.; Atkinson, S.; Przysowa, J.; Lojkowska, E.; Williams, P.; Jafra, S. Inactivation of AHLs by Ochrobactrum sp. A44 depends on the activity of a novel class of AHL acylase. Environ. Microbiol. Rep. 2011, 3, 59–68. [Google Scholar] [CrossRef]
  56. Valera, M.J.; Mas, A.; Streit, W.R.; Mateo, E. GqqA, a novel protein in Komagataeibacter europaeus involved in bacterial quorum quenching and cellulose formation. Microb. Cell Fact. 2016, 15, 88. [Google Scholar] [CrossRef]
  57. Werner, N.; Petersen, K.; Vollstedt, C.; Garcia, P.P.; Chow, J.; Ferrer, M.; Fernandez-Lopez, L.; Falke, S.; Perbandt, M.; Hinrichs, W.; et al. The Komagataeibacter europaeus GqqA is the prototype of a novel bifunctional N-Acyl-homoserine lactone acylase with prephenate dehydratase activity. Sci. Rep. 2021, 11, 12255. [Google Scholar] [CrossRef]
  58. Shepherd, R.W.; Lindow, S.E. Two dissimilar N-acyl-homoserine lactone acylases of Pseudomonas syringae influence colony and biofilm morphology. Appl. Environ. Microbiol. 2009, 75, 45–53. [Google Scholar] [CrossRef]
  59. Kusada, H.; Tamaki, H.; Kamagata, Y.; Hanada, S.; Kimura, N. A Novel Quorum-Quenching N-Acylhomoserine Lactone Acylase from Acidovorax sp. Strain MR-S7 Mediates Antibiotic Resistance. Appl. Environ. Microbiol. 2017, 83, e00080-17. [Google Scholar] [CrossRef]
  60. Sio, C.F.; Otten, L.G.; Cool, R.H.; Diggle, S.P.; Braun, P.G.; Bos, R.; Daykin, M.; Cámara, M.; Williams, P.; Quax, W.J. Quorum quenching by an N-acyl-homoserine lactone acylase from Pseudomonas aeruginosa PAO1. Infect. Immun. 2006, 74, 1673–1682. [Google Scholar] [CrossRef]
  61. Utari, P.D.; Setroikromo, R.; Melgert, B.N.; Quax, W.J. PvdQ Quorum Quenching Acylase Attenuates Pseudomonas aeruginosa Virulence in a Mouse Model of Pulmonary Infection. Front. Cell Infect. Microbiol. 2018, 8, 119. [Google Scholar] [CrossRef] [PubMed]
  62. Bokhove, M.; Nadal Jimenez, P.; Quax, W.J.; Dijkstra, B.W. The quorum-quenching N-acyl homoserine lactone acylase PvdQ is an Ntn-hydrolase with an unusual substrate-binding pocket. Proc. Natl. Acad. Sci. USA 2010, 107, 686–691. [Google Scholar] [CrossRef]
  63. Huang, J.J.; Petersen, A.; Whiteley, M.; Leadbetter, J.R. Identification of QuiP, the product of gene PA1032, as the second acyl-homoserine lactone acylase of Pseudomonas aeruginosa PAO1. Appl. Environ. Microbiol. 2006, 72, 1190–1197. [Google Scholar] [CrossRef] [PubMed]
  64. Philem, P.D.; Yadav, Y.; Vellore Sunder, A.; Ghosh, D.; Prabhune, A.; Ramasamy, S. Structural and enzymatic analysis of a dimeric cholylglycine hydrolase like acylase active on N-acyl homoserine lactones. Biochimie 2020, 177, 108–116. [Google Scholar] [CrossRef]
  65. Lord, D.M.; Baran, A.U.; Wood, T.K.; Peti, W.; Page, R. BdcA, a protein important for Escherichia coli biofilm dispersal, is a short-chain dehydrogenase/reductase that binds specifically to NADPH. PLoS ONE 2014, 9, e105751. [Google Scholar] [CrossRef] [PubMed]
  66. Bijtenhoorn, P.; Schipper, C.; Hornung, C.; Quitschau, M.; Grond, S.; Weiland, N.; Streit, W.R. BpiB05, a novel metagenome-derived hydrolase acting on N-acylhomoserine lactones. J. Biotechnol. 2011, 155, 86–94. [Google Scholar] [CrossRef] [PubMed]
  67. Bijtenhoorn, P.; Mayerhofer, H.; Müller-Dieckmann, J.; Utpatel, C.; Schipper, C.; Hornung, C.; Szesny, M.; Grond, S.; Thürmer, A.; Brzuszkiewicz, E.; et al. A novel metagenomic short-chain dehydrogenase/reductase attenuates Pseudomonas aeruginosa biofilm formation and virulence on Caenorhabditis elegans. PLoS ONE 2011, 6, e26278. [Google Scholar] [CrossRef]
  68. Chowdhary, P.K.; Keshavan, N.; Nguyen, H.Q.; Peterson, J.A.; González, J.E.; Haines, D.C. Bacillus megaterium CYP102A1 oxidation of acyl homoserine lactones and acyl homoserines. Biochemistry 2007, 46, 14429–14437. [Google Scholar] [CrossRef]
  69. Uroz, S.; Chhabra, S.R.; Cámara, M.; Williams, P.; Oger, P.; Dessaux, Y. N-Acylhomoserine lactone quorum-sensing molecules are modified and degraded by Rhodococcus erythropolis W2 by both amidolytic and novel oxidoreductase activities. Microbiology 2005, 151, 3313–3322. [Google Scholar] [CrossRef]
  70. Weiland-Bräuer, N.; Kisch, M.J.; Pinnow, N.; Liese, A.; Schmitz, R.A. Highly Effective Inhibition of Biofilm Formation by the First Metagenome-Derived AI-2 Quenching Enzyme. Front. Microbiol. 2016, 7, 1098. [Google Scholar] [CrossRef]
  71. Dong, Y.H.; Wang, L.H.; Xu, J.L.; Zhang, H.B.; Zhang, X.F.; Zhang, L.H. Quenching quorum-sensing-dependent bacterial infection by an N-acyl homoserine lactonase. Nature 2001, 411, 813–817. [Google Scholar] [CrossRef] [PubMed]
  72. Dong, Y.H.; Xu, J.L.; Li, X.Z.; Zhang, L.H. AiiA, an enzyme that inactivates the acylhomoserine lactone quorum-sensing signal and attenuates the virulence of Erwinia carotovora. Proc. Natl. Acad. Sci. USA 2000, 97, 3526–3531. [Google Scholar] [CrossRef] [PubMed]
  73. Bergonzi, C.; Schwab, M.; Naik, T.; Daudé, D.; Chabrière, E.; Elias, M. Structural and Biochemical Characterization of AaL, a Quorum Quenching Lactonase with Unusual Kinetic Properties. Sci. Rep. 2018, 8, 11262. [Google Scholar] [CrossRef] [PubMed]
  74. Kalia, V.C.; Raju, S.C.; Purohit, H.J. Genomic analysis reveals versatile organisms for quorum quenching enzymes: Acyl-homoserine lactone-acylase and -lactonase. Open Microbiol. J. 2011, 5, 1–13. [Google Scholar] [CrossRef]
  75. Leadbetter, J.R.; Greenberg, E.P. Metabolism of acyl-homoserine lactone quorum-sensing signals by Variovorax paradoxus. J. Bacteriol. 2000, 182, 6921–6926. [Google Scholar] [CrossRef]
  76. Yu, M.; Liu, N.; Zhao, Y.; Zhang, X. Quorum quenching enzymes of marine bacteria and implication in aquaculture. J. Ocean Univ. 2018, 48, 34–43. [Google Scholar]
  77. Marques, J.C.; Lamosa, P.; Russell, C.; Ventura, R.; Maycock, C.; Semmelhack, M.F.; Miller, S.T.; Xavier, K.B. Processing the interspecies quorum-sensing signal autoinducer-2 (AI-2): Characterization of phospho-(S)-4,5-dihydroxy-2,3-pentanedione isomerization by LsrG protein. J. Biol. Chem. 2011, 286, 18331–18343. [Google Scholar] [CrossRef]
  78. Xavier, K.B.; Miller, S.T.; Lu, W.; Kim, J.H.; Rabinowitz, J.; Pelczer, I.; Semmelhack, M.F.; Bassler, B.L. Phosphorylation and processing of the quorum-sensing molecule autoinducer-2 in enteric bacteria. ACS Chem. Biol. 2007, 2, 128–136. [Google Scholar] [CrossRef]
  79. Gatta, V.; Ilina, P.; Porter, A.; McElroy, S.; Tammela, P. Targeting Quorum Sensing: High-Throughput Screening to Identify Novel LsrK Inhibitors. Int. J. Mol. Sci. 2019, 20, 3112. [Google Scholar] [CrossRef]
  80. Stotani, S.; Gatta, V.; Medarametla, P.; Padmanaban, M.; Karawajczyk, A.; Giordanetto, F.; Tammela, P.; Laitinen, T.; Poso, A.; Tzalis, D.; et al. DPD-Inspired Discovery of Novel LsrK Kinase Inhibitors: An Opportunity To Fight Antimicrobial Resistance. J. Med. Chem. 2019, 62, 2720–2737. [Google Scholar] [CrossRef]
  81. Medarametla, P.; Gatta, V.; Kajander, T.; Laitinen, T.; Tammela, P.; Poso, A. Structure-Based Virtual Screening of LsrK Kinase Inhibitors to Target Quorum Sensing. Chem. Med. Chem. 2018, 13, 2400–2407. [Google Scholar] [CrossRef] [PubMed]
  82. Yu, L.; Li, W.; Zhang, M.; Cui, Y.; Chen, X.; Ni, J.; Yu, L.; Shang, F.; Xue, T. Imidazole decreases the ampicillin resistance of an Escherichia coli strain isolated from a cow with mastitis by inhibiting the function of autoinducer 2. J. Dairy Sci. 2018, 101, 3356–3362. [Google Scholar] [CrossRef] [PubMed]
  83. Hernando-Amado, S.; Alcalde-Rico, M.; Gil-Gil, T.; Valverde, J.R.; Martínez, J.L. Naringenin Inhibition of the Pseudomonas aeruginosa Quorum Sensing Response Is Based on Its Time-Dependent Competition with N-(3-Oxo-dodecanoyl)-L-homoserine Lactone for LasR Binding. Front. Mol. Biosci. 2020, 7, 25. [Google Scholar] [CrossRef] [PubMed]
  84. Paczkowski, J.E.; Mukherjee, S.; McCready, A.R.; Cong, J.P.; Aquino, C.J.; Kim, H.; Henke, B.R.; Smith, C.D.; Bassler, B.L. Flavonoids Suppress Pseudomonas aeruginosa Virulence through Allosteric Inhibition of Quorum-sensing Receptors. J. Biol. Chem. 2017, 292, 4064–4076. [Google Scholar] [CrossRef] [PubMed]
  85. Woods, K.E.; Akhter, S.; Rodriguez, B.; Townsend, K.A.; Smith, N.; Smith, B.; Wambua, A.; Craddock, V.; Abisado-Duque, R.G.; Santa, E.E.; et al. Characterization of natural product inhibitors of quorum sensing in Pseudomonas aeruginosa reveals competitive inhibition of RhlR by ortho-vanillin. bioRxiv 2024. [Google Scholar] [CrossRef]
  86. Khayat, M.T.; Abbas, H.A.; Ibrahim, T.S.; Khayyat, A.N.; Alharbi, M.; Darwish, K.M.; Elhady, S.S.; Khafagy, E.S.; Safo, M.K.; Hegazy, W.A.H. Anti-Quorum Sensing Activities of Gliptins against Pseudomonas aeruginosa and Staphylococcus aureus. Biomedicines 2022, 10, 1169. [Google Scholar] [CrossRef] [PubMed]
  87. Abbas, H.A.; Shaldam, M.A.; Eldamasi, D. Curtailing Quorum Sensing in Pseudomonas aeruginosa by Sitagliptin. Curr. Microbiol. 2020, 77, 1051–1060. [Google Scholar] [CrossRef]
  88. Dwivedi, D.; Singh, V. Effects of the natural compounds embelin and piperine on the biofilm-producing property of Streptococcus mutans. J. Tradit. Complement. Med. 2015, 6, 57–61. [Google Scholar] [CrossRef] [PubMed]
  89. Khan, F.; Lee, J.W.; Javaid, A.; Park, S.K.; Kim, Y.M. Inhibition of biofilm and virulence properties of Pseudomonas aeruginosa by sub-inhibitory concentrations of aminoglycosides. Microb. Pathog. 2020, 146, 104249. [Google Scholar] [CrossRef]
  90. Manson, D.E.; Ananiev, G.E.; Guo, S.; Ericksen, S.S.; Santa, E.E.; Blackwell, H.E. Abiotic Small Molecule Inhibitors and Activators of the LasR Quorum Sensing Receptor in Pseudomonas aeruginosa with Potencies Comparable or Surpassing N-Acyl Homoserine Lactones. ACS Infect. Dis. 2024, 10, 1212–1221. [Google Scholar] [CrossRef]
  91. Reverchon, S.; Chantegrel, B.; Deshayes, C.; Doutheau, A.; Cotte-Pattat, N. New synthetic analogues of N-acyl homoserine lactones as agonists or antagonists of transcriptional regulators involved in bacterial quorum sensing. Bioorg. Med. Chem. Lett. 2002, 12, 1153–1157. [Google Scholar] [CrossRef] [PubMed]
  92. Schaefer, A.L.; Hanzelka, B.L.; Eberhard, A.; Greenberg, E.P. Quorum sensing in Vibrio fischeri: Probing autoinducer-LuxR interactions with autoinducer analogs. J. Bacteriol. 1996, 178, 2897–2901. [Google Scholar] [CrossRef] [PubMed]
  93. Soltane, R.; Alasiri, A.; Taha, M.N.; Abd El-Aleam, R.H.; Alghamdi, K.S.; Ghareeb, M.A.; Keshek, D.E.; Cardoso, S.M.; Sayed, A.M. Norlobaridone Inhibits Quorum Sensing-Dependent Biofilm Formation and Some Virulence Factors in Pseudomonas aeruginosa by Disrupting Its Transcriptional Activator Protein LasR Dimerization. Biomolecules 2023, 13, 1573. [Google Scholar] [CrossRef]
  94. Chaieb, K.; Kouidhi, B.; Hosawi, S.B.; Baothman, O.A.S.; Zamzami, M.A.; Altayeb, H.N. Computational screening of natural compounds as putative quorum sensing inhibitors targeting drug resistance bacteria: Molecular docking and molecular dynamics simulations. Comput. Biol. Med. 2022, 145, 105517. [Google Scholar] [CrossRef] [PubMed]
  95. Serrano-Aguirre, L.; Velasco-Bucheli, R.; García-Álvarez, B.; Saborido, A.; Arroyo, M.; de la Mata, I. Novel Bifunctional Acylase from Actinoplanes utahensis: A Versatile Enzyme to Synthesize Antimicrobial Compounds and Use in Quorum Quenching Processes. Antibiotics 2021, 10, 922. [Google Scholar] [CrossRef] [PubMed]
  96. Gopu, V.; Meena, C.K.; Shetty, P.H. Quercetin Influences Quorum Sensing in Food Borne Bacteria: In-Vitro and In-Silico Evidence. PLoS ONE 2015, 10, e0134684. [Google Scholar] [CrossRef] [PubMed]
  97. Geske, G.D.; O’Neill, J.C.; Miller, D.M.; Wezeman, R.J.; Mattmann, M.E.; Lin, Q.; Blackwell, H.E. Comparative analyses of N-acylated homoserine lactones reveal unique structural features that dictate their ability to activate or inhibit quorum sensing. ChemBioChem 2008, 9, 389–400. [Google Scholar] [CrossRef]
  98. Garner, A.L.; Park, J.; Zakhari, J.S.; Lowery, C.A.; Struss, A.K.; Sawada, D.; Kaufmann, G.F.; Janda, K.D. A multivalent probe for AI-2 quorum-sensing receptors. J. Am. Chem. Soc. 2011, 133, 15934–15937. [Google Scholar] [CrossRef]
  99. Peng, H.; Cheng, Y.; Ni, N.; Li, M.; Choudhary, G.; Chou, H.T.; Lu, C.D.; Tai, P.C.; Wang, B. Synthesis and evaluation of new antagonists of bacterial quorum sensing in Vibrio harveyi. Chem. Med. Chem. 2009, 4, 1457–1468. [Google Scholar] [CrossRef]
  100. Bernabè, G.; Dal Pra, M.; Ronca, V.; Pauletto, A.; Marzaro, G.; Saluzzo, F.; Stefani, A.; Artusi, I.; De Filippis, V.; Ferlin, M.G.; et al. A Novel Aza-Derivative Inhibits agr Quorum Sensing Signaling and Synergizes Methicillin-Resistant Staphylococcus aureus to Clindamycin. Front. Microbiol. 2021, 12, 610859. [Google Scholar] [CrossRef]
  101. Mahdally, N.H.; George, R.F.; Kashef, M.T.; Al-Ghobashy, M.; Murad, F.E.; Attia, A.S. Staquorsin: A Novel Staphylococcus aureus Agr-Mediated Quorum Sensing Inhibitor Impairing Virulence in vivo without Notable Resistance Development. Front. Microbiol. 2021, 12, 700494. [Google Scholar] [CrossRef] [PubMed]
  102. Palaniappan, B.; Solomon, A.P. Targeting AgrA quorum sensing regulator by bumetanide attenuates virulence in Staphylococcus aureus—A drug repurposing approach. Life Sci. 2021, 273, 119306. [Google Scholar] [CrossRef] [PubMed]
  103. Zhu, J.; Beaber, J.W.; Moré, M.I.; Fuqua, C.; Eberhard, A.; Winans, S.C. Analogs of the autoinducer 3-oxooctanoyl-homoserine lactone strongly inhibit activity of the TraR protein of Agrobacterium tumefaciens. J. Bacteriol. 1998, 180, 5398–5405. [Google Scholar] [CrossRef] [PubMed]
  104. Oh, M.H.; Choi, C.H. Role of LuxIR Homologue AnoIR in Acinetobacter nosocomialis and the Effect of Virstatin on the Expression of anoR Gene. J. Microbiol. Biotechnol. 2015, 25, 1390–1400. [Google Scholar] [CrossRef] [PubMed]
  105. El-Shaer, S.; Shaaban, M.; Barwa, R.; Hassan, R. Control of quorum sensing and virulence factors of Pseudomonas aeruginosa using phenylalanine arginyl β-naphthylamide. J. Med. Microbiol. 2016, 65, 1194–1204. [Google Scholar] [CrossRef]
  106. Escobar-Muciño, E.; Arenas-Hernández, M.M.P.; Luna-Guevara, M.L. Mechanisms of Inhibition of Quorum Sensing as an Alternative for the Control of E. coli and Salmonella. Microorganisms 2022, 10, 884. [Google Scholar] [CrossRef] [PubMed]
  107. Xie, Y.; Chen, J.M.; Wang, B.; Peng, A.Y.; Mao, Z.W.; Xia, W. Inhibition of Quorum-Sensing Regulator from Pseudomonas aeruginosa Using a Flavone Derivative. Molecules 2022, 27, 2439. [Google Scholar] [CrossRef] [PubMed]
  108. Atchan, A.P.N.; Monthe, O.C.; Tchamgoue, A.D.; Singh, Y.; Shivashankara, S.T.; Selvi, M.K.; Agbor, G.A.; Magni, P.; Piazza, S.; Manjappara, U.V.; et al. Anti-Inflammatory, Antioxidant Activities, and Phytochemical Characterization of Edible Plants Exerting Synergistic Effects in Human Gastric Epithelial Cells. Antioxidants 2023, 12, 591. [Google Scholar] [CrossRef]
  109. Dincheva, I.; Badjakov, I.; Galunska, B. New Insights into the Research of Bioactive Compounds from Plant Origins with Nutraceutical and Pharmaceutical Potential. Plants 2023, 12, 258. [Google Scholar] [CrossRef]
  110. Sharifi-Rad, J.; Sureda, A.; Tenore, G.C.; Daglia, M.; Sharifi-Rad, M.; Valussi, M.; Tundis, R.; Sharifi-Rad, M.; Loizzo, M.R.; Ademiluyi, A.O.; et al. Biological Activities of Essential Oils: From Plant Chemoecology to Traditional Healing Systems. Molecules 2017, 22, 70. [Google Scholar] [CrossRef]
  111. Aswathanarayan, J.B.; Vittal, R.R. Inhibition of biofilm formation and quorum sensing mediated phenotypes by berberine in Pseudomonas aeruginosa and Salmonella typhimurium. RSC Adv. 2018, 8, 36133–36141. [Google Scholar] [CrossRef]
  112. Qian, W.D.; Huang, J.; Zhang, J.N.; Li, X.C.; Kong, Y.; Wang, T.; Li, Y.D. Antimicrobial and Antibiofilm Activities and Mechanism of Action of Chelerythrine Against Carbapenem-Resistant Serratia marcescens In Vitro. Microb. Drug Resist. 2021, 27, 1105–1116. [Google Scholar] [CrossRef] [PubMed]
  113. Tajani, A.S.; Amiri Tehranizadeh, Z.; Pourmohammad, A.; Pourmohammad, A.; Iranshahi, M.; Farhadi, F.; Soheili, V.; Fazly Bazzaz, B.S. Anti-quorum sensing and antibiofilm activity of coumarin derivatives against Pseudomonas aeruginosa PAO1: Insights from in vitro and in silico studies. Iran. J. Basic Med. Sci. 2023, 26, 445–452. [Google Scholar]
  114. He, Z.; Jiang, W.; Jiang, Y.; Dong, J.; Song, Z.; Xu, J.; Zhou, W. Anti-biofilm activities of coumarin as quorum sensing inhibitor for Porphyromonas gingivalis. J. Oral Microbiol. 2022, 14, 2055523. [Google Scholar] [CrossRef] [PubMed]
  115. Majik, M.S.; Naik, D.; Bhat, C.; Tilve, S.; Tilvi, S.; D’Souza, L. Synthesis of (R)-norbgugaine and its potential as quorum sensing inhibitor against Pseudomonas aeruginosa. Bioorg. Med. Chem. Lett. 2013, 23, 2353–2356. [Google Scholar] [CrossRef]
  116. Janeczko, M.; Masłyk, M.; Kubiński, K.; Golczyk, H. Emodin, a natural inhibitor of protein kinase CK2, suppresses growth, hyphal development, and biofilm formation of Candida albicans. Yeast 2017, 34, 253–265. [Google Scholar] [CrossRef] [PubMed]
  117. Yan, X.; Gu, S.; Shi, Y.; Cui, X.; Wen, S.; Ge, J. The effect of emodin on Staphylococcus aureus strains in planktonic form and biofilm formation in vitro. Arch. Microbiol. 2017, 199, 1267–1275. [Google Scholar] [CrossRef] [PubMed]
  118. Vikram, A.; Jayaprakasha, G.K.; Uckoo, R.M.; Patil, B.S. Inhibition of Escherichia coli O157:H7 motility and biofilm by β-sitosterol glucoside. Biochim. Biophys. Acta 2013, 1830, 5219–5228. [Google Scholar] [CrossRef] [PubMed]
  119. Vikram, A.; Jesudhasan, P.R.; Jayaprakasha, G.K.; Pillai, S.D.; Patil, B.S. Citrus limonoids interfere with Vibrio harveyi cell-cell signalling and biofilm formation by modulating the response regulator LuxO. Microbiology 2011, 157, 99–110. [Google Scholar] [CrossRef]
  120. Quave, C.L.; Estévez-Carmona, M.; Compadre, C.M.; Hobby, G.; Hendrickson, H.; Beenken, K.E.; Smeltzer, M.S. Ellagic acid derivatives from Rubus ulmifolius inhibit Staphylococcus aureus biofilm formation and improve response to antibiotics. PLoS ONE 2012, 7, e28737. [Google Scholar] [CrossRef]
  121. Konozy, E.H.E.; Osman, M.E.M.; Dirar, A.I.; Ghartey-Kwansah, G. Plant lectins: A new antimicrobial frontier. Biomed. Pharmacother. 2022, 155, 113735. [Google Scholar] [CrossRef] [PubMed]
  122. Procópio, T.F.; de Siqueira Patriota, L.L.; de Moura, M.C.; da Silva, P.M.; de Oliveira, A.P.; do Nascimento Carvalho, L.V.; de Albuquerque Lima, T.; Soares, T.; da Silva, T.D.; Breitenbach Barroso Coelho, L.C.; et al. CasuL: A new lectin isolated from Calliandra surinamensis leaf pinnulae with cytotoxicity to cancer cells, antimicrobial activity and antibiofilm effect. Int. J. Biol. Macromol. 2017, 98, 419–429. [Google Scholar] [CrossRef]
  123. Jin, X.; Lee, Y.J.; Hong, S.H. Canavalia ensiformis-derived lectin inhibits biofilm formation of enterohemorrhagic Escherichia coli and Listeria monocytogenes. J. Appl. Microbiol. 2019, 126, 300–310. [Google Scholar] [CrossRef]
  124. Cavalcante, T.T.; Anderson Matias da Rocha, B.; Alves Carneiro, V.; Vassiliepe Sousa Arruda, F.; Fernandes do Nascimento, A.S.; Cardoso Sá, N.; do Nascimento, K.S.; Sousa Cavada, B.; Holanda Teixeira, E. Effect of lectins from Diocleinae subtribe against oral Streptococci. Molecules 2011, 16, 3530–3543. [Google Scholar] [CrossRef]
  125. Silva, P.M.; Napoleão, T.H.; Silva, L.C.; Fortes, D.T.; Lima, T.A.; Zingali, R.B.; Pontual, E.V.; Araújo, J.M.; Medeiros, P.L.; Rodrigues, C.G. The juicy sarcotesta of Punica granatum contains a lectin that affects growth, survival as well as adherence and invasive capacities of human pathogenic bacteria. J. Funct. Foods 2016, 27, 695–702. [Google Scholar] [CrossRef]
  126. Emad, F.; Khalafalah, A.K.; El Sayed, M.A.; Mohamed, A.H.; Stadler, M.; Helaly, S.E. Three new polyacetylene glycosides (PAGs) from the aerial part of Launaea capitata (Asteraceae) with anti-biofilm activity against Staphylococcus aureus. Fitoterapia 2020, 143, 104548. [Google Scholar] [CrossRef] [PubMed]
  127. Carneiro, V.A.; Santos, H.S.; Arruda, F.V.; Bandeira, P.N.; Albuquerque, M.R.; Pereira, M.O.; Henriques, M.; Cavada, B.S.; Teixeira, E.H. Casbane diterpene as a promising natural antimicrobial agent against biofilm-associated infections. Molecules 2010, 16, 190–201. [Google Scholar] [CrossRef] [PubMed]
  128. Murata, R.M.; Branco-de-Almeida, L.S.; Franco, E.M.; Yatsuda, R.; dos Santos, M.H.; de Alencar, S.M.; Koo, H.; Rosalen, P.L. Inhibition of Streptococcus mutans biofilm accumulation and development of dental caries in vivo by 7-epiclusianone and fluoride. Biofouling 2010, 26, 865–872. [Google Scholar] [CrossRef] [PubMed]
  129. Raorane, C.J.; Lee, J.H.; Kim, Y.G.; Rajasekharan, S.K.; García-Contreras, R.; Lee, J. Antibiofilm and Antivirulence Efficacies of Flavonoids and Curcumin Against Acinetobacter baumannii. Front. Microbiol. 2019, 10, 990. [Google Scholar] [CrossRef]
  130. Luo, J.; Dong, B.; Wang, K.; Cai, S.; Liu, T.; Cheng, X.; Lei, D.; Chen, Y.; Li, Y.; Kong, J.; et al. Baicalin inhibits biofilm formation, attenuates the quorum sensing-controlled virulence and enhances Pseudomonas aeruginosa clearance in a mouse peritoneal implant infection model. PLoS ONE 2017, 12, e0176883. [Google Scholar] [CrossRef]
  131. Das, M.C.; Sandhu, P.; Gupta, P.; Rudrapaul, P.; De, U.C.; Tribedi, P.; Akhter, Y.; Bhattacharjee, S. Attenuation of Pseudomonas aeruginosa biofilm formation by Vitexin: A combinatorial study with azithromycin and gentamicin. Sci. Rep. 2016, 6, 23347. [Google Scholar] [CrossRef] [PubMed]
  132. Rayan, M.; Abu Lafi, S.; Falah, M.; Kacergius, T.; Kirkliauskiene, A.; Gabe, V.; Rayan, A. Alkyl Gallates as Potential Antibiofilm Agents: A Review. Molecules 2023, 28, 1751. [Google Scholar] [CrossRef] [PubMed]
  133. Rudin, L.; Bornstein, M.M.; Shyp, V. Inhibition of biofilm formation and virulence factors of cariogenic oral pathogen Streptococcus mutans by natural flavonoid phloretin. J. Oral Microbiol. 2023, 15, 2230711. [Google Scholar] [CrossRef]
  134. Tapia-Rodriguez, M.R.; Bernal-Mercado, A.T.; Gutierrez-Pacheco, M.M.; Vazquez-Armenta, F.J.; Hernandez-Mendoza, A.; Gonzalez-Aguilar, G.A.; Martinez-Tellez, M.A.; Nazzaro, F.; Ayala-Zavala, J.F. Virulence of Pseudomonas aeruginosa exposed to carvacrol: Alterations of the Quorum sensing at enzymatic and gene levels. J. Cell Commun. Signal. 2019, 13, 531–537. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, W.J.; Liu, C.C.; Li, Y.T.; Li, M.Q.; Fu, Y.T.; Li, X.C.; Qian, W.D. Antifungal and Antibiofilm In Vitro Activities of Ursolic Acid on Cryptococcus neoformans. Curr. Microbiol. 2022, 79, 293. [Google Scholar] [CrossRef] [PubMed]
  136. Ali, F.; Sangwan, P.L.; Koul, S.; Pandey, A.; Bani, S.; Abdullah, S.T.; Sharma, P.R.; Kitchlu, S.; Khan, I.A. 4-epi-Pimaric acid: A phytomolecule as a potent antibacterial and anti-biofilm agent for oral cavity pathogens. Eur. J. Clin. Microbiol. Infect. Dis. 2012, 31, 149–159. [Google Scholar] [CrossRef] [PubMed]
  137. Ušjak, L.; Stojković, D.; Carević, T.; Milutinović, V.; Soković, M.; Niketić, M.; Petrović, S. Chemical Analysis and Investigation of Antimicrobial and Antibiofilm Activities of Prangos trifida (Apiaceae). Antibiotics 2024, 13, 41. [Google Scholar] [CrossRef] [PubMed]
  138. D’Aquila, P.; Sena, G.; Crudo, M.; Passarino, G.; Bellizzi, D. Effect of Essential Oils of Apiaceae, Lamiaceae, Lauraceae, Myrtaceae, and Rutaceae Family Plants on Growth, Biofilm Formation, and Quorum Sensing in Chromobacterium violaceum, Pseudomonas aeruginosa, and Enterococcus faecalis. Microorganisms 2023, 11, 1150. [Google Scholar] [CrossRef]
  139. Alibi, S.; Ben Selma, W.; Ramos-Vivas, J.; Smach, M.A.; Touati, R.; Boukadida, J.; Navas, J.; Ben Mansour, H. Anti-oxidant, antibacterial, anti-biofilm, and anti-quorum sensing activities of four essential oils against multidrug-resistant bacterial clinical isolates. Curr. Res. Transl. Med. 2020, 68, 59–66. [Google Scholar] [CrossRef]
  140. Condò, C.; Anacarso, I.; Sabia, C.; Iseppi, R.; Anfelli, I.; Forti, L.; de Niederhäusern, S.; Bondi, M.; Messi, P. Antimicrobial activity of spices essential oils and its effectiveness on mature biofilms of human pathogens. Nat. Prod. Res. 2020, 34, 567–574. [Google Scholar] [CrossRef]
  141. Sharifi, A.; Mohammadzadeh, A.; Zahraei Salehi, T.; Mahmoodi, P. Antibacterial, antibiofilm and antiquorum sensing effects of Thymus daenensis and Satureja hortensis essential oils against Staphylococcus aureus isolates. J. Appl. Microbiol. 2018, 124, 379–388. [Google Scholar] [CrossRef] [PubMed]
  142. Porfírio, E.M.; Melo, H.M.; Pereira, A.M.G.; Cavalcante, T.T.A.; Gomes, G.A.; de Carvalho, M.G.; Costa, R.A.; Júnior, F.E.A.C. In Vitro Antibacterial and Antibiofilm Activity of Lippia alba Essential Oil, Citral, and Carvone against Staphylococcus aureus. Sci. World J. 2017, 2017, 4962707. [Google Scholar] [CrossRef]
  143. Vieira, M.; Bessa, L.J.; Martins, M.R.; Arantes, S.; Teixeira, A.P.S.; Mendes, Â.; Martins da Costa, P.; Belo, A.D.F. Chemical Composition, Antibacterial, Antibiofilm and Synergistic Properties of Essential Oils from Eucalyptus globulus Labill. and Seven Mediterranean Aromatic Plants. Chem. Biodivers. 2017, 14, e1700006. [Google Scholar] [CrossRef] [PubMed]
  144. Lyu, X.; Lee, J.; Chen, W.N. Potential Natural Food Preservatives and Their Sustainable Production in Yeast: Terpenoids and Polyphenols. J. Agric. Food Chem. 2019, 67, 4397–4417. [Google Scholar] [CrossRef] [PubMed]
  145. Han, Y.; Sun, Z.; Chen, W. Antimicrobial susceptibility and antibacterial mechanism of limonene against Listeria monocytogenes. Molecules 2020, 25, 33. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, X.; Cai, J.; Chen, H.; Zhong, Q.; Hou, Y.; Chen, W.; Chen, W. Antibacterial activity and mechanism of linalool against Pseudomonas aeruginosa. Microb. Pathog. 2020, 141, 103980. [Google Scholar] [CrossRef]
  147. D’Aquila, P.; Paparazzo, E.; Crudo, M.; Bonacci, S.; Procopio, A.; Passarino, G.; Bellizzi, D. Antibacterial Activity and Epigenetic Remodeling of Essential Oils from Calabrian Aromatic Plants. Nutrients 2022, 14, 391. [Google Scholar] [CrossRef] [PubMed]
  148. Behzadnia, A.; Moosavi-Nasab, M.; Oliyaei, N. Anti-biofilm activity of marine algae-derived bioactive compounds. Front. Microbiol. 2024, 15, 1270174. [Google Scholar] [CrossRef]
  149. Al-Tmimi, S.L.; Al-Rrubaai, G.; Zaki, N.H. Antibiofilm activity of intracellular extracts of Westiellopsis prolifica isolated from local environment in Baghdad. JGPT 2018, 10, 281–288. [Google Scholar]
  150. El-Sapagh, S.; El-Shenody, R.; Pereira, L.; Elshobary, M. Unveiling the Potential of Algal Extracts as Promising Antibacterial and Antibiofilm Agents against Multidrug-Resistant Pseudomonas aeruginosa: In Vitro and In Silico Studies including Molecular Docking. Plants 2023, 12, 3324. [Google Scholar] [CrossRef]
  151. Nithya, C.; LewisOscar, F.; Kanaga, S.; Kavitha, R.; Bakkiyaraj, D.; Arunkumar, M.; Alharbi, N.S.; Chinnathambi, A.; Ali Alharbi, S.; Thajuddin, N. Biofilm inhibitory potential of Chlamydomonas sp. extract against Pseudomonas aeruginosa. J. Algal Biomass Utln. 2014, 5, 74–81. [Google Scholar]
  152. El Zawawy, N.A.; El-Shenody, R.A.; Ali, S.S.; El-Shetehy, M. A novel study on the inhibitory effect of marine macroalgal extracts on hyphal growth and biofilm formation of candidemia isolates. Sci. Rep. 2020, 10, 9339. [Google Scholar] [CrossRef]
  153. Cepas, V.; Gutiérrez-Del-Río, I.; López, Y.; Redondo-Blanco, S.; Gabasa, Y.; Iglesias, M.J.; Soengas, R.; Fernández-Lorenzo, A.; López-Ibáñez, S.; Villar, C.J.; et al. Microalgae and Cyanobacteria Strains as Producers of Lipids with Antibacterial and Antibiofilm Activity. Mar. Drugs 2021, 19, 675. [Google Scholar] [CrossRef]
  154. Besednova, N.N.; Andryukov, B.G.; Zaporozhets, T.S.; Kryzhanovsky, S.P.; Kuznetsova, T.A.; Fedyanina, L.N.; Makarenkova, I.D.; Zvyagintseva, T.N. Algae Polyphenolic Compounds and Modern Antibacterial Strategies: Current Achievements and Immediate Prospects. Biomedicines 2020, 8, 342. [Google Scholar] [CrossRef]
  155. Prasath, K.G.; Tharani, H.; Kumar, M.S.; Pandian, S.K. Palmitic Acid Inhibits the Virulence Factors of Candida tropicalis: Biofilms, Cell Surface Hydrophobicity, Ergosterol Biosynthesis, and Enzymatic Activity. Front. Microbiol. 2020, 11, 864. [Google Scholar] [CrossRef]
  156. Boutin, R.; Munnier, E.; Renaudeau, N.; Girardot, M.; Pinault, M.; Chevalier, S.; Chourpa, I.; Clément-Larosière, B.; Imbertet, C.; Boudesocque-Delaye, L. Spirulina platensis sustainable lipid extracts in alginate-based nanocarriers: An algal approach against biofilms. Algal Res. 2019, 37, 160–168. [Google Scholar] [CrossRef]
  157. Sampathkumar, S.J.; Srivastava, P.; Ramachandran, S.; Sivashanmugam, K.; Gothandam, K.M. Lutein: A potential antibiofilm and antiquorum sensing molecule from green microalga Chlorella pyrenoidosa. Microb. Pathog. 2019, 135, 103658. [Google Scholar] [CrossRef]
  158. Vishwakarma, J.; Vavilala, S.L. Evaluating the antibacterial and antibiofilm potential of sulphated polysaccharides extracted from green algae Chlamydomonas reinhardtii. J. Appl. Microbiol. 2019, 127, 1004–1017. [Google Scholar] [CrossRef]
  159. Jun, J.Y.; Jung, M.J.; Jeong, I.H.; Yamazaki, K.; Kawai, Y.; Kim, B.M. Antimicrobial and Antibiofilm Activities of Sulfated Polysaccharides from Marine Algae against Dental Plaque Bacteria. Mar. Drugs 2018, 16, 301. [Google Scholar] [CrossRef]
  160. Powell, L.C.; Pritchard, M.F.; Ferguson, E.L.; Powell, K.A.; Patel, S.U.; Rye, P.D.; Sakellakou, S.M.; Buurma, N.J.; Brilliant, C.D.; Copping, J.M.; et al. Targeted disruption of the extracellular polymeric network of Pseudomonas aeruginosa biofilms by alginate oligosaccharides. npj Biofilms Microbiomes 2018, 4, 13. [Google Scholar] [CrossRef]
  161. Gökalsın, B.; Aksoydan, B.; Erman, B.; Sesal, N.C. Reducing Virulence and Biofilm of Pseudomonas aeruginosa by Potential Quorum Sensing Inhibitor Carotenoid: Zeaxanthin. Microb. Ecol. 2017, 74, 466–473. [Google Scholar] [CrossRef] [PubMed]
  162. Melander, R.J.; Basak, A.K.; Melander, C. Natural products as inspiration for the development of bacterial antibiofilm agents. Nat. Prod. Rep. 2020, 37, 1454–1477. [Google Scholar] [CrossRef]
  163. You, J.; Xue, X.; Cao, L.; Lu, X.; Wang, J.; Zhang, L.; Zhou, S. Inhibition of Vibrio biofilm formation by a marine actinomycete strain A66. Appl. Microbiol. Biotechnol. 2007, 76, 1137–1144. [Google Scholar] [CrossRef] [PubMed]
  164. Oloke, J.K. Activity pattern of natural and synthetic antibacterial agents among hospital isolates. Microbios 2000, 102, 175–181. [Google Scholar] [PubMed]
  165. Eguchi, Y.; Kubo, N.; Matsunaga, H.; Igarashi, M.; Utsumi, R. Development of an antivirulence drug against Streptococcus mutans: Repression of biofilm formation, acid tolerance, and competence by a histidine kinase inhibitor, walkmycin C. Antimicrob. Agents Chemother. 2011, 55, 1475–1484. [Google Scholar] [CrossRef]
  166. Vasilchenko, A.S.; Julian, W.T.; Lapchinskaya, O.A.; Katrukha, G.S.; Sadykova, V.S.; Rogozhin, E.A. A Novel Peptide Antibiotic Produced by Streptomyces roseoflavus Strain INA-Ac-5812 With Directed Activity Against Gram-Positive Bacteria. Front. Microbiol. 2020, 11, 556063. [Google Scholar] [CrossRef] [PubMed]
  167. Bauermeister, A.; Pereira, F.; Grilo, I.R.; Godinho, C.C.; Paulino, M.; Almeida, V.; Gobbo-Neto, L.; Prieto-Davó, A.; Sobral, R.G.; Lopes, N.P.; et al. Intra-clade metabolomic profiling of MAR4 Streptomyces from the Macaronesia Atlantic region reveals a source of anti-biofilm metabolites. Environ. Microbiol. 2019, 21, 1099–1112. [Google Scholar] [CrossRef]
  168. Driche, E.H.; Sabaou, N.; Bijani, C.; Zitouni, A.; Pont, F.; Mathieu, F.; Badji, B. Streptomyces sp. AT37 isolated from a Saharan soil produces a furanone derivative active against multidrug-resistant Staphylococcus aureus. World J. Microbiol. Biotechnol. 2017, 33, 105. [Google Scholar] [CrossRef] [PubMed]
  169. Oja, T.; San Martin Galindo, P.; Taguchi, T.; Manner, S.; Vuorela, P.M.; Ichinose, K.; Metsä-Ketelä, M.; Fallarero, A. Effective Antibiofilm Polyketides against Staphylococcus aureus from the Pyranonaphthoquinone Biosynthetic Pathways of Streptomyces Species. Antimicrob. Agents Chemother. 2015, 59, 6046–6052. [Google Scholar] [CrossRef]
  170. Suzuki, N.; Ohtaguro, N.; Yoshida, Y.; Hirai, M.; Matsuo, H.; Yamada, Y.; Imamura, N.; Tsuchiya, T. A Compound Inhibits Biofilm Formation of Staphylococcus aureus from Streptomyces. Biol. Pharm. Bull. 2015, 38, 889–892. [Google Scholar] [CrossRef]
  171. Guillén-Navarro, D.; González-Vázquez, R.; León-Ávila, G.; Giono-Cerezo, S. Quorum Quenching with a Diffusible Signal Factor Analog in Stenotrophomonas maltophilia. Pathogens 2023, 12, 1448. [Google Scholar] [CrossRef]
  172. Zhou, L.; Zhang, L.H.; Cámara, M.; He, Y.W. The DSF Family of Quorum Sensing Signals: Diversity, Biosynthesis, and Turnover. Trends Microbiol. 2017, 25, 293–303. [Google Scholar] [CrossRef] [PubMed]
  173. Marques, C.N.; Davies, D.G.; Sauer, K. Control of Biofilms with the Fatty Acid Signaling Molecule cis-2-Decenoic Acid. Pharmaceuticals 2015, 8, 816–835. [Google Scholar] [CrossRef] [PubMed]
  174. Apel, C.; Barg, A.; Rheinberg, A.; Conrads, G.; Wagner-Döbler, I. Dental composite materials containing carolacton inhibit biofilm growth of Streptococcus mutans. Dent. Mater. 2013, 29, 1188–1199. [Google Scholar] [CrossRef] [PubMed]
  175. Abouelhassan, Y.; Garrison, A.T.; Yang, H.; Chávez-Riveros, A.; Burch, G.M.; Huigens, R.W., 3rd. Recent Progress in Natural-Product-Inspired Programs Aimed to Address Antibiotic Resistance and Tolerance. J. Med. Chem. 2019, 62, 7618–7642. [Google Scholar] [CrossRef] [PubMed]
  176. Gayathiri, E.; Prakash, P.; Karmegam, N.; Varjani, S.; Awasthi, M.K.; Ravindran, B. Biosurfactants: Potential and eco-friendly material for sustainable agriculture and environmental safety—A review. Agronomy 2022, 12, 662. [Google Scholar] [CrossRef]
  177. Paraszkiewicz, K.; Moryl, M.; Płaza, G.; Bhagat, D.K.; Satpute, S.; Bernat, P. Surfactants of microbial origin as antibiofilm agents. Int. J. Environ. Health Res. 2021, 31, 401–420. [Google Scholar] [CrossRef] [PubMed]
  178. Abdel-Aziz, M.M.; Al-Omar, M.S.; Mohammed, H.A.; Emam, T.M. In Vitro and Ex Vivo Antibiofilm Activity of a Lipopeptide Biosurfactant Produced by the Entomopathogenic Beauveria bassiana Strain against Microsporum canis. Microorganisms 2020, 8, 232. [Google Scholar] [CrossRef]
  179. Abdollahi, S.; Tofighi, Z.; Babaee, T.; Shamsi, M.; Rahimzadeh, G.; Rezvanifar, H.; Saeidi, E.; Mohajeri Amiri, M.; Saffari Ashtiani, Y.; Samadi, N. Evaluation of Anti-oxidant and Anti-biofilm Activities of Biogenic Surfactants Derived from Bacillus amyloliquefaciens and Pseudomonas aeruginosa. Iran. J. Pharm. Res. 2020, 19, 115–126. [Google Scholar]
  180. Shakerifard, P.; Gancel, F.; Jacques, P.; Faille, C. Effect of different Bacillus subtilis lipopeptides on surface hydrophobicity and adhesion of Bacillus cereus 98/4 spores to stainless steel and Teflon. Biofouling 2009, 25, 533–541. [Google Scholar] [CrossRef]
  181. Pamp, S.J.; Tolker-Nielsen, T. Multiple roles of biosurfactants in structural biofilm development by Pseudomonas aeruginosa. J. Bacteriol. 2007, 189, 2531–2539. [Google Scholar] [CrossRef] [PubMed]
  182. Yan, X.; Gu, S.; Cui, X.; Shi, Y.; Wen, S.; Chen, H.; Ge, J. Antimicrobial, anti-adhesive and anti-biofilm potential of biosurfactants isolated from Pediococcus acidilactici and Lactobacillus plantarum against Staphylococcus aureus CMCC26003. Microb. Pathog. 2019, 127, 12–20. [Google Scholar] [CrossRef] [PubMed]
  183. Batoni, G.; Maisetta, G.; Esin, S. Antimicrobial peptides and their interaction with biofilms of medically relevant bacteria. Biochim. Biophys. Acta 2016, 1858, 1044–1060. [Google Scholar] [CrossRef] [PubMed]
  184. Pletzer, D.; Coleman, S.R.; Hancock, R.E. Anti-biofilm peptides as a new weapon in antimicrobial warfare. Curr. Opin. Microbiol. 2016, 33, 35–40. [Google Scholar] [CrossRef] [PubMed]
  185. Di Somma, A.; Moretta, A.; Canè, C.; Cirillo, A.; Duilio, A. Antimicrobial and Antibiofilm Peptides. Biomolecules 2020, 10, 652. [Google Scholar] [CrossRef]
  186. Yasir, M.; Willcox, M.D.P.; Dutta, D. Action of Antimicrobial Peptides against Bacterial Biofilms. Materials 2018, 11, 2468. [Google Scholar] [CrossRef] [PubMed]
  187. Arslan, S.Y.; Leung, K.P.; Wu, C.D. The effect of lactoferrin on oral bacterial attachment. Oral Microbiol. Immunol. 2009, 24, 411–416. [Google Scholar] [CrossRef]
  188. Wang, H.Y.; Lin, L.; Tan, L.S.; Yu, H.Y.; Cheng, J.W.; Pan, Y.P. Molecular pathways underlying inhibitory effect of antimicrobial peptide Nal-P-113 on bacteria biofilms formation of Porphyromonas gingivalis W83 by DNA microarray. BMC Microbiol. 2017, 17, 37. [Google Scholar] [CrossRef] [PubMed]
  189. Okuda, K.; Zendo, T.; Sugimoto, S.; Iwase, T.; Tajima, A.; Yamada, S.; Sonomoto, K.; Mizunoe, Y. Effects of bacteriocins on methicillin-resistant Staphylococcus aureus biofilm. Antimicrob. Agents Chemother. 2013, 57, 5572–5579. [Google Scholar] [CrossRef]
  190. Luca, V.; Stringaro, A.; Colone, M.; Pini, A.; Mangoni, M.L. Esculentin(1-21), an amphibian skin membrane-active peptide with potent activity on both planktonic and biofilm cells of the bacterial pathogen Pseudomonas aeruginosa. Cell Mol. Life Sci. 2013, 70, 2773–2786. [Google Scholar] [CrossRef]
  191. Nagant, C.; Pitts, B.; Stewart, P.S.; Feng, Y.; Savage, P.B.; Dehaye, J.P. Study of the effect of antimicrobial peptide mimic, CSA-13, on an established biofilm formed by Pseudomonas aeruginosa. Microbiologyopen 2013, 2, 318–325. [Google Scholar] [CrossRef] [PubMed]
  192. Jacobsen, A.S.; Jenssen, H. Human cathelicidin LL-37 prevents bacterial biofilm formation. Future Med. Chem. 2012, 4, 1587–1599. [Google Scholar] [CrossRef]
  193. de la Fuente-Núñez, C.; Korolik, V.; Bains, M.; Nguyen, U.; Breidenstein, E.B.; Horsman, S.; Lewenza, S.; Burrows, L.; Hancock, R.E. Inhibition of bacterial biofilm formation and swarming motility by a small synthetic cationic peptide. Antimicrob. Agents Chemother. 2012, 56, 2696–2704. [Google Scholar] [CrossRef]
  194. Zhu, C.; Tan, H.; Cheng, T.; Shen, H.; Shao, J.; Guo, Y.; Shi, S.; Zhang, X. Human β-defensin 3 inhibits antibiotic-resistant Staphylococcus biofilm formation. J. Surg. Res. 2013, 183, 204–213. [Google Scholar] [CrossRef]
  195. Ansari, J.M.; Abraham, N.M.; Massaro, J.; Murphy, K.; Smith-Carpenter, J.; Fikrig, E. Anti-Biofilm Activity of a Self-Aggregating Peptide against Streptococcus mutans. Front. Microbiol. 2017, 8, 488. [Google Scholar] [CrossRef]
  196. Gordya, N.; Yakovlev, A.; Kruglikova, A.; Tulin, D.; Potolitsina, E.; Suborova, T.; Bordo, D.; Rosano, C.; Chernysh, S. Natural antimicrobial peptide complexes in the fighting of antibiotic resistant biofilms: Calliphora vicina medicinal maggots. PLoS ONE 2017, 12, e0173559. [Google Scholar] [CrossRef]
  197. Brancatisano, F.L.; Maisetta, G.; Di Luca, M.; Esin, S.; Bottai, D.; Bizzarri, R.; Campa, M.; Batoni, G. Inhibitory effect of the human liver-derived antimicrobial peptide hepcidin 20 on biofilms of polysaccharide intercellular adhesin (PIA)-positive and PIA-negative strains of Staphylococcus epidermidis. Biofouling 2014, 30, 435–446. [Google Scholar] [CrossRef]
  198. Mathur, H.; Field, D.; Rea, M.C.; Cotter, P.D.; Hill, C.; Ross, R.P. Fighting biofilms with lantibiotics and other groups of bacteriocins. NPJ Biofilms Microbiomes 2018, 4, 9. [Google Scholar] [CrossRef]
  199. Chopra, L.; Singh, G.; Choudhary, V.; Sahoo, D.K. Sonorensin: An antimicrobial peptide, belonging to the heterocycloanthracin subfamily of bacteriocins, from a new marine isolate, Bacillus sonorensis MT93. Appl. Environ. Microbiol. 2014, 80, 2981–2990. [Google Scholar] [CrossRef]
  200. Jiang, Y.H.; Xin, W.G.; Zhang, Q.L.; Lin, L.B.; Deng, X.Y. A Novel Bacteriocin against Shigella flexneri from Lactiplantibacillus plantarum Isolated from Tilapia Intestine: Purification, Antibacterial Properties and Antibiofilm Activity. Front. Microbiol. 2022, 12, 779315. [Google Scholar] [CrossRef]
  201. Yoon, J.W.; Kang, S.S. In Vitro Antibiofilm and Anti-Inflammatory Properties of Bacteriocins Produced by Pediococcus acidilactici Against Enterococcus faecalis. Foodborne Pathog. Dis. 2020, 17, 764–771. [Google Scholar] [CrossRef] [PubMed]
  202. Ray Mohapatra, A.; Jeevaratnam, K. Inhibiting bacterial colonization on catheters: Antibacterial and antibiofilm activities of bacteriocins from Lactobacillus plantarum SJ33. J. Glob. Antimicrob. Resist. 2019, 19, 85–92. [Google Scholar] [CrossRef] [PubMed]
  203. Vashistha, A.; Sharma, N.; Nanaji, Y.; Kumar, D.; Singh, G.; Barnwal, R.P.; Yadav, A.K. Quorum sensing inhibitors as Thera-peutics: Bacterial biofilm inhibition. Bioorg. Chem. 2023, 136, 106551. [Google Scholar] [CrossRef] [PubMed]
  204. Zhan, Y.; Hu, H.; Yu, Y.; Chen, C.; Zhang, J.; Jarnda, K.V.; Ding, P. Therapeutic strategies for drug-resistant Pseudomonas aeruginosa: Metal and metal oxide nanoparticles. J. Biomed. Mater. Res. A 2024. [Google Scholar] [CrossRef] [PubMed]
  205. Guglielmelli, A.; D’Aquila, P.; Palermo, G.; Dell’Aglio, M.; Passarino, G.; Strangi, G.; Bellizzi, D. Role of the Human Serum Albumin Protein Corona in the Antimicrobial and Photothermal Activity of Metallic Nanoparticles against Escherichia coli Bacteria. ACS Omega 2023, 8, 31333–31343. [Google Scholar] [CrossRef] [PubMed]
  206. Shaikh, S.; Nazam, N.; Rizvi, S.M.D.; Ahmad, K.; Baig, M.H.; Lee, E.J.; Choi, I. Mechanistic Insights into the Antimicrobial Actions of Metallic Nanoparticles and Their Implications for Multidrug Resistance. Int. J. Mol. Sci. 2019, 20, 2468. [Google Scholar] [CrossRef] [PubMed]
  207. Niemirowicz, K.; Swiecicka, I.; Wilczewska, A.Z.; Misztalewska, I.; Kalska-Szostko, B.; Bienias, K.; Bucki, R.; Car, H. Gold-functionalized magnetic nanoparticles restrict growth of Pseudomonas aeruginosa. Int. J. Nanomed. 2014, 9, 2217–2224. [Google Scholar]
  208. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal nanoparticles: Understanding the mechanisms behind antibacterial activity. J. Nanobiotechnol. 2017, 15, 65. [Google Scholar] [CrossRef]
  209. Robino, L.; Scavone, P. Nanotechnology in biofilm prevention. Future Microbiol. 2020, 15, 377–379. [Google Scholar] [CrossRef]
  210. Qayyum, S.; Khan, A.U. Nanoparticles: Vs. biofilms: A battle against another paradigm of antibiotic resistance. MedChemComm 2016, 7, 1479–1498. [Google Scholar] [CrossRef]
  211. Slomberg, D.L.; Lu, Y.; Broadnax, A.D.; Hunter, R.A.; Carpenter, A.W.; Schoenfisch, M.H. Role of size and shape on biofilm eradication for nitric oxide-releasing silica nanoparticles. ACS Appl. Mater. Interfaces 2013, 5, 9322–9329. [Google Scholar] [CrossRef] [PubMed]
  212. Lellouche, J.; Friedman, A.; Lellouche, J.P.; Gedanken, A.; Banin, E. Improved antibacterial and antibiofilm activity of magnesium fluoride nanoparticles obtained by water-based ultrasound chemistry. Nanomedicine 2012, 8, 702–711. [Google Scholar] [CrossRef] [PubMed]
  213. Li, P.; Chen, X.; Shen, Y.; Li, H.; Zou, Y.; Yuan, G.; Hu, P.; Hu, H. Mucus penetration enhanced lipid polymer nanoparticles improve the eradication rate of Helicobacter pylori biofilm. J. Control. Release 2019, 300, 52–63. [Google Scholar] [CrossRef]
  214. Li, X.; Yeh, Y.C.; Giri, K.; Mout, R.; Landis, R.F.; Prakash, Y.S.; Rotello, V.M. Control of nanoparticle penetration into biofilms through surface design. Chem. Commun. 2015, 51, 282–285. [Google Scholar] [CrossRef] [PubMed]
  215. Fulaz, S.; Vitale, S.; Quinn, L.; Casey, E. Nanoparticle-Biofilm Interactions: The Role of the EPS Matrix. Trends Microbiol. 2019, 27, 915–926. [Google Scholar] [CrossRef]
  216. Omran, B.A.; Tseng, B.S.; Baek, K.H. Nanocomposites against Pseudomonas aeruginosa biofilms: Recent advances, challenges, and future prospects. Microbiol. Res. 2024, 282, 127656. [Google Scholar] [CrossRef]
  217. Kapoor, D.U.; Patel, R.J.; Gaur, M.; Parikh, S.; Prajapati, B.G. Metallic and metal oxide nanoparticles in treating Pseudomonas aeruginosa infections. J. Drug Deliv. Sci. Technol. 2024, 91, 105290. [Google Scholar] [CrossRef]
  218. Baptista, P.V.; McCusker, M.P.; Carvalho, A.; Ferreira, D.A.; Mohan, N.M.; Martins, M.; Fernandes, A.R. Nano-Strategies to Fight Multidrug Resistant Bacteria—“A Battle of the Titans”. Front. Microbiol. 2018, 9, 1441. [Google Scholar] [CrossRef]
  219. El-Deeb, N.M.; Abo-Eleneen, M.A.; Al-Madboly, L.A.; Sharaf, M.M.; Othman, S.S.; Ibrahim, O.M.; Mubarak, M.S. Biogenically Synthesized Polysaccharides-Capped Silver Nanoparticles: Immunomodulatory and Antibacterial Potentialities Against Resistant Pseudomonas aeruginosa. Front. Bioeng. Biotechnol. 2020, 8, 643. [Google Scholar] [CrossRef]
  220. LewisOscar, F.; Nithya, C.; Vismaya, S.; Arunkumar, M.; Pugazhendhi, A.; Nguyen-Tri, P.; Alharbi, S.A.; Alharbi, N.S.; Thajuddin, N. In vitro analysis of green fabricated silver nanoparticles (AgNPs) against Pseudomonas aeruginosa PA14 biofilm formation, their application on urinary catheter. Prog. Org. Coat. 2021, 151, 106058. [Google Scholar] [CrossRef]
  221. Hu, D.; Li, H.; Wang, B.; Ye, Z.; Lei, W.; Jia, F.; Jin, Q.; Ren, K.F.; Ji, J. Surface-Adaptive Gold Nanoparticles with Effective Adherence and Enhanced Photothermal Ablation of Methicillin-Resistant Staphylococcus aureus Biofilm. ACS Nano 2017, 11, 9330–9339. [Google Scholar] [CrossRef] [PubMed]
  222. Rajkumari, J.; Busi, S.; Vasu, A.C.; Reddy, P. Facile green synthesis of baicalein fabricated gold nanoparticles and their antibiofilm activity against Pseudomonas aeruginosa PAO1. Microb. Pathog. 2017, 107, 261–269. [Google Scholar] [CrossRef] [PubMed]
  223. Lee, J.H.; Kim, Y.G.; Cho, M.H.; Lee, J. ZnO nanoparticles inhibit Pseudomonas aeruginosa biofilm formation and virulence factor production. Microbiol. Res. 2014, 169, 888–896. [Google Scholar] [CrossRef]
  224. Javanbakht, T.; Laurent, S.; Stanicki, D.; Wilkinson, K.J. Relating the Surface Properties of Superparamagnetic Iron Oxide Nanoparticles (SPIONs) to Their Bactericidal Effect towards a Biofilm of Streptococcus mutans. PLoS ONE 2016, 11, e0154445. [Google Scholar] [CrossRef] [PubMed]
  225. Li, W.; Wei, W.; Wu, X.; Zhao, Y.; Dai, H. The antibacterial and antibiofilm activities of mesoporous hollow Fe3O4 nanoparticles in an alternating magnetic field. Biomater. Sci. 2020, 8, 4492–4507. [Google Scholar] [CrossRef]
  226. Subbiahdoss, G.; Sharifi, S.; Grijpma, D.W.; Laurent, S.; van der Mei, H.C.; Mahmoudi, M.; Busscher, H.J. Magnetic targeting of surface-modified superparamagnetic iron oxide nanoparticles yields antibacterial efficacy against biofilms of gentamicin-resistant staphylococci. Acta Biomater. 2012, 8, 2047–2055. [Google Scholar] [CrossRef] [PubMed]
  227. Auger, S.; Henry, C.; Péchoux, C.; Suman, S.; Lejal, N.; Bertho, N.; Larcher, T.; Stankic, S.; Vidic, J. Exploring multiple effects of Zn0.15Mg0.85O nanoparticles on Bacillus subtilis and macrophages. Sci. Rep. 2018, 8, 12276. [Google Scholar] [CrossRef]
  228. Eshed, M.; Lellouche, J.; Gedanken, A.; Banin, E. A Zn-doped CuO nanocomposite shows enhanced antibiofilm and antibacterial activities against Streptococcus mutans compared to nanosized CuO. Adv. Funct. Mater. 2014, 24, 1382–1390. [Google Scholar] [CrossRef]
  229. Ho, K.K.; Cole, N.; Chen, R.; Willcox, M.D.; Rice, S.A.; Kumar, N. Immobilization of antibacterial dihydropyrrol-2-ones on functional polymer supports to prevent bacterial infections in vivo. Antimicrob. Agents Chemother. 2012, 56, 1138–1141. [Google Scholar] [CrossRef]
  230. Roncari Rocha, G.; Sims, K.R., Jr.; Xiao, B.; Klein, M.I.; Benoit, D.S.W. Nanoparticle carrier co-delivery of complementary antibiofilm drugs abrogates dual species cariogenic biofilm formation in vitro. J. Oral Microbiol. 2021, 14, 1997230. [Google Scholar] [CrossRef]
  231. Vinoj, G.; Pati, R.; Sonawane, A.; Vaseeharan, B. In vitro cytotoxic effects of gold nanoparticles coated with functional acyl homoserine lactone lactonase protein from Bacillus licheniformis and their antibiofilm activity against Proteus species. Antimicrob. Agents Chemother. 2015, 59, 763–771. [Google Scholar] [CrossRef]
  232. Lee, H.W.; Kharel, S.; Loo, S.C.J. Lipid-Coated Hybrid Nanoparticles for Enhanced Bacterial Biofilm Penetration and Antibiofilm Efficacy. ACS Omega 2022, 7, 35814–35824. [Google Scholar] [CrossRef]
  233. Maurizi, L.; Forte, J.; Ammendolia, M.G.; Hanieh, P.N.; Conte, A.L.; Relucenti, M.; Donfrancesco, O.; Ricci, C.; Rinaldi, F.; Marianecci, C.; et al. Effect of Ciprofloxacin-Loaded Niosomes on Escherichia coli and Staphylococcus aureus Biofilm Formation. Pharmaceutics 2022, 14, 2662. [Google Scholar] [CrossRef]
  234. Lambadi, P.R.; Sharma, T.K.; Kumar, P.; Vasnani, P.; Thalluri, S.M.; Bisht, N.; Pathania, R.; Navani, N.K. Facile biofunctionalization of silver nanoparticles for enhanced antibacterial properties, endotoxin removal, and biofilm control. Int. J. Nanomed. 2015, 10, 2155–2171. [Google Scholar]
  235. Ho, D.K.; Murgia, X.; De Rossi, C.; Christmann, R.; Hüfner de Mello Martins, A.G.; Koch, M.; Andreas, A.; Herrmann, J.; Müller, R.; Empting, M.; et al. Squalenyl Hydrogen Sulfate Nanoparticles for Simultaneous Delivery of Tobramycin and an Alkylquinolone Quorum Sensing Inhibitor Enable the Eradication of P. aeruginosa Biofilm Infections. Angew. Chem. Int. Ed. Engl. 2020, 59, 10292–10296. [Google Scholar] [CrossRef]
  236. Ivanova, K.; Ivanova, A.; Hoyo, J.; Pérez-Rafael, S.; Tzanov, T. Nano-Formulation Endows Quorum Quenching Enzyme-Antibiotic Hybrids with Improved Antibacterial and Antibiofilm Activities against Pseudomonas aeruginosa. Int. J. Mol. Sci. 2022, 23, 7632. [Google Scholar] [CrossRef]
  237. Charkowski, A.O. The Changing Face of Bacterial Soft-Rot Diseases. Annu. Rev. Phytopathol. 2018, 56, 269–288. [Google Scholar] [CrossRef]
  238. Rodríguez, M.; Torres, M.; Blanco, L.; Béjar, V.; Sampedro, I.; Llamas, I. Plant growth-promoting activity and quorum quenching-mediated biocontrol of bacterial phytopathogens by Pseudomonas segetis strain P6. Sci. Rep. 2020, 10, 4121. [Google Scholar] [CrossRef]
  239. Dong, Y.H.; Zhang, X.F.; Xu, J.L.; Zhang, L.H. Insecticidal Bacillus thuringiensis silences Erwinia carotovora virulence by a new form of microbial antagonism, signal interference. Appl. Environ. Microbiol. 2004, 70, 954–960. [Google Scholar] [CrossRef]
  240. Zhang, W.; Fan, X.; Li, J.; Ye, T.; Mishra, S.; Zhang, L.; Chen, S. Exploration of the Quorum-Quenching Mechanism in Pseudomonas nitroreducens W-7 and Its Potential to Attenuate the Virulence of Dickeya zeae EC1. Front. Microbiol. 2021, 12, 694161. [Google Scholar] [CrossRef]
  241. Yu, H.; Chen, W.J.; Bhatt, K.; Zhou, Z.; Zhu, X.; Liu, S.; He, J.; Zhang, L.H.; Chen, S.; Wang, H.; et al. A novel bacterial strain Burkholderia sp. F25 capable of degrading diffusible signal factor signal shows strong biocontrol potential. Front. Plant Sci. 2022, 13, 1071693. [Google Scholar] [CrossRef] [PubMed]
  242. Ye, T.; Zhang, W.; Feng, Z.; Fan, X.; Xu, X.; Mishra, S.; Zhang, L.; Chen, S. Characterization of a Novel Quorum-Quenching Bacterial Strain, Burkholderia anthina HN-8, and Its Biocontrol Potential against Black Rot Disease Caused by Xanthomonas campestris pv. campestris. Microorganisms 2020, 8, 1485. [Google Scholar] [CrossRef]
  243. Ye, T.; Zhou, T.; Fan, X.; Bhatt, P.; Zhang, L.; Chen, S. Acinetobacter lactucae Strain QL-1, a Novel Quorum Quenching Candidate Against Bacterial Pathogen Xanthomonas campestris pv. campestris. Front. Microbiol. 2019, 10, 2867. [Google Scholar] [CrossRef] [PubMed]
  244. D’Angelo-Picard, C.; Chapelle, E.; Ratet, P.; Faure, D.; Dessaux, Y. Transgenic plants expressing the quorum quenching lactonase AttM do not significantly alter root-associated bacterial populations. Res. Microbiol. 2011, 162, 951–958. [Google Scholar] [CrossRef]
  245. Chu, W.; Zhou, S.; Zhu, W.; Zhuang, X. Quorum quenching bacteria Bacillus sp. QSI-1 protect zebrafish (Danio rerio) from Aeromonas hydrophila infection. Sci. Rep. 2014, 4, 5446. [Google Scholar] [CrossRef] [PubMed]
  246. Vinoj, G.; Vaseeharan, B.; Thomas, S.; Spiers, A.J.; Shanthi, S. Quorum-quenching activity of the AHL-lactonase from Bacillus licheniformis DAHB1 inhibits Vibrio biofilm formation in vitro and reduces shrimp intestinal colonisation and mortality. Mar. Biotechnol. 2014, 16, 707–715. [Google Scholar] [CrossRef]
  247. Monzón-Atienza, L.; Bravo, J.; Torrecillas, S.; Gómez-Mercader, A.; Montero, D.; Ramos-Vivas, J.; Galindo-Villegas, J.; Acosta, F. An In-Depth Study on the Inhibition of Quorum Sensing by Bacillus velezensis D-18: Its Significant Impact on Vibrio Biofilm Formation in Aquaculture. Microorganisms 2024, 12, 890. [Google Scholar] [CrossRef]
  248. Chen, B.; Peng, M.; Tong, W.; Zhang, Q.; Song, Z. The Quorum Quenching Bacterium Bacillus licheniformis T-1 Protects Zebrafish against Aeromonas hydrophila Infection. Probiotics Antimicrob. Proteins 2020, 12, 160–171. [Google Scholar] [CrossRef]
  249. Ghanei-Motlagh, R.; Mohammadian, T.; Gharibi, D.; Menanteau-Ledouble, S.; Mahmoudi, E.; Khosravi, M.; Zarea, M.; El-Matbouli, M. Quorum Quenching Properties and Probiotic Potentials of Intestinal Associated Bacteria in Asian Sea Bass Lates calcarifer. Mar. Drugs 2019, 18, 23. [Google Scholar] [CrossRef]
  250. Zhou, S.; Yu, Z.; Chu, W. Effect of quorum-quenching bacterium Bacillus sp. QSI-1 on protein profiles and extracellular enzymatic activities of Aeromonas hydrophila YJ-1. BMC Microbiol. 2019, 19, 135. [Google Scholar] [CrossRef]
  251. Maddela, N.R.; Abiodun, A.S.; Zhang, S.; Prasad, R. Biofouling in Membrane Bioreactors-Mitigation and Current Status: A Review. Appl. Biochem. Biotechnol. 2023, 195, 5643–5668. [Google Scholar] [CrossRef] [PubMed]
  252. Ergön-Can, T.; Köse-Mutlu, B.; Koyuncu, I.; Lee, C.H. Biofouling control based on bacterial quorum quenching with a new application: Rotary microbial carrier frame. J. Membr. Sci. 2017, 525, 116–124. [Google Scholar] [CrossRef]
  253. Lade, H.; Paul, D.; Kweon, J.H. Quorum quenching mediated approaches for control of membrane biofouling. Int. J. Biol. Sci. 2014, 10, 550–565. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of the different levels at which Quorum Sensing acts in regulating biofilm formation and compounds effective in its inhibition. AI: autoinducer. Figure made using Biorender.com.
Figure 1. Schematic diagram of the different levels at which Quorum Sensing acts in regulating biofilm formation and compounds effective in its inhibition. AI: autoinducer. Figure made using Biorender.com.
Antibiotics 13 00619 g001
Table 1. Quorum Quenching enzymes, their substrates, and their relative mechanisms of inhibition of biofilm.
Table 1. Quorum Quenching enzymes, their substrates, and their relative mechanisms of inhibition of biofilm.
Enzyme CategoryNameSourceSubstratesEffectReference
LactonasesAiiABacillus sp. Priestia aryabhattaiC4-, C6-, C8-, C10-HSL; 3OC4-, 3OC6-, 3OC8-, 3OC12-HSL; 3-OH-C4-HSLInhibition of biofilm and production of pyocyanin, rhamnolipid, and exopolysaccarides in P. aeruginosa, Vibrio cholerae, and S. aureus.[36,37]
AiiBAgrobacterium tumefaciensC4-, C6-, C7-, C8-, C10-HSL; 3OC6-, 3OC8-HSLReduction in the bacterial virulence in Erwinia carotovora.[38]
AiiKLactobacillus casei, Kurthia huakuiC10-HSLAttenuation of swimming motility, virulence factor production, and biofilm formation in Aeromonas hydrophila and P. aeruginosa[39,40]
Aii810Bacteria from Mao-tofu3OC12-HSLInhibition of virulence and biofilm in P. aeruginosa.[41]
bpiB01, B04, B07Bacteria from soil samples3OC8-HSL; 3OC6-, 3OC8-, 3OC12-HSLBiofilm inhibition in P. aeruginosa.[42]
DlhR, QsdR1Rhizobium sp.3OC8-HSLInhibition of biofilm and other QS-dependent processes in P. aeruginosa, Chromobacterium violaceum, and A. tumefaciens.[43]
NDGeobacillus kaustophilus3-OH-C10-HSL, 3-OH-C12-HSLDistruption of biofilm on multiple strains of A. baumannii.[44]
MCPMycobacterium aviumC6-, C7-, C8-, C10-, C12-HSLThe enzyme is a metal-dependent N-acylhomoserine lactonase[45]
MomLMuricauda oleariaC6-HSLAttenuation of the virulence of P. aeruginosa.[46]
SsoPox-1Sulfolobus solfataricusC4-, C6-, C8-, C12-HSL; 3OC6-, 3OC8-, 3OC10-, 3OC12-HSLAttenuation of QS signaling, virulence factor production, and biofilm formation in vitro.[47]
PON1-3Homo sapiensC7-, C12-, C14-HSL; 3OC6-, 3OC10-, 3OC12-HSLInhibition of biofilm formation and extracellular virulence factors in P. aeruginosa.[48]
YtnPStenotrophomonas maltophilia3O-C12-HSLInhibition of biofilm formation, induction of biofilm decomposition, and reducetion of extracellular virulence factors production.[49]
AcylasesAacRalstonia solanacearumC7-, C8-, C10-HSL; 3OC8-HSLInhibition of QS mechanism in C. violaceum.[50]
AhlDArthrobacter sp.OHHL, OHL and OdDHReduction in the AHL amount and pectate lyase activity in Erwinia carotovora.[51]
AhlMStreptomyces sp.C6-, C8-, C10-HSL; 3OC6-, 3OC8-, 3OC12-HSLDecrease of the production of virulence factors, including elastase, total protease, and LasA, in P. aeruginosa.[52]
aibPBrucella melitensis3OC8-, 3OC12-HSLDecrease of endogenous AHL accumulation within Brucella melitensis.[53]
AiiDRalstonia sp.OC6-, OC8-, OC10-, OC12-HSLQQ activity in P. aeruginosa achieved through an Acyl-homoserine lactone acylase hydrolyses the AHL amide, releasing homoserine lactone and the corresponding fatty acid.[54]
AiiOOchrobactrum sp. A443OC4-14-HSLInhibition of the virulence in Pectobacterium carotovorum.[55]
GqqAKomagataeibacter europaeusC8-, C10-, C12-HSInhibiton of biofilm formation.[56,57]
HacA, BPseudomonas syringaeC6-, C8-, C10-, C12-HSL; OC8-, OC10-, OC12-, OC14-HSLInhibition of biofilm formation, and colony morphology.[58]
MacQAcidovorax sp.C6-, C8-, C10-, C12-HSL; 3OC8-HSL; OC6-, OC8-, OC10-, OC12-, OC14-HSLInterference with the QS system.[59]
PA2385P. aeruginosa3COC12-HSLQQ activity achieved through an N-acyl-homoserine lactone acylase[60]
PvdQP. aeruginosaC8-, C10-, C12-HSL; OC12-HSLAttenuation of P. aeruginosa virulence.[61,62]
QuiPP. aeruginosaC8-, C10-, C12-HSL; OC12-HSLGene expression regulation of QS systems in P. aeruginosa.[63]
Slac1, 2Shewanella loihica-PV43OC10-HSLExhibition of in vitro QQ activity.[64]
OxidoreductasesBdcAE. coli3OC6-HSLDehydrogenase/reductase activity.[65]
BpiB05Bacteria from soil sample3-oxo-C8-HSLExhibition of in vitro QQ activity.[66]
BpiB09Bacteria from soil sample3-oxo-C8-HSLReduction in pyocyanin production, decreased motility, poor biofilm formation in P. aeruginosa.[67]
CYP102A1Bacillus megateriumC12-20-HSL (ω-1, ω-2, ω-3 hydroxylated)In vitro oxidation of AHLs and their lactonolysis products acyl homoserines.[68]
NDRhodococcus erythropolis3-oxo-AHL, N-(3-oxo-6-phenylhexanoyl) homoserine lactone, 3-oxododecanamide, n-(3-oxododecanoyl)-L-homoserine lactoneExhibition of AHL oxidoreductase and amidolytic activities.[69]
QQ-2Bacteria from water sample3-oxo-C6-HSL; 3,4,4-trihydroxy-2-pentanone-5-phosphateReduction of AHL and AI-2 in QS-inactive hydroxy-derivatives and inhibits biofilm formation in Klebsiella oxytoca and Klebsiella pneumoniae.[70]
Table 2. Quorum Quenching natural compounds and their relative molecular effects on biofilm.
Table 2. Quorum Quenching natural compounds and their relative molecular effects on biofilm.
Phytochemical ClassCompoundsNatural SourceMolecular EffectReference
AlkaloidsBerberineCoptis, Hydrastis and Berberis genusDownregulation of QS-related genes.[111]
ChelerythrineChelidonium majus L.Disruption of membrane integrity, inhibition of biofilm components.[112]
FurocoumarinsGrapefruitsInhibition of QS signalling; Biofilm dispersion.[113,114]
NorbgugaineArisarum vulgareInhibition of cell motility and of adhesion.[115]
AnthraquinonesEmodinRheum palmatum L.Suppression growth, hyphal development, and biofilm of Candida albicans by targeting cellular kinase signalling.[116]
Polygonum cuspidatum, Rheum palmatumDownregulation of biofilm-forming related genes in S. aureus.[117]
Glycosidesβ-sitosterol glucoside; Isolimonic acid; IchanginCitrus speciesInhibition of the biofilm and motility through the repression of flagellar master operon flhDC Interferece with cell–cell signalling and biofilm formation by the modulation of luxO expression.[118,119]
Ellagic acidRubus ulmifoliusInhibition of the biofilm and increase antibiotic susceptibility.[120]
LectinsAGLAmaranthus gangeticusInhibition of biofilm formation.[121]
CasuLCalliandra surinamensiInhibition of biofilm formation.[122]
ConACanavalia ensiformisReduction in bacterial adhesion.[123]
ConBolCanavalia bolivianaReduction in bacterial adhesion.[124]
ConMCanavalia maritimaReduction in bacterial adhesion.[124]
PgTeLPunica granatumReduction in bacterial adhesion.[125]
PolyacetylenesGlicosides (PAGs)Launaea capitata [126]
PolyphenolsCasbane diterpeneCroton nepetaefoliusInhibition of biofilm and reduction of bacterial adhesion.[127]
7-epiclusianoneRheedia brasiliensisInhibition of biofilm and reduction of bacterial adhesion.[128]
Flavonoids (baicalin, catechin, curcumin, kaempherol, luteolin, proaminophyllin, quercetin, vitexin)Scutellaria baicalensis and Vitex speciesDecrease in the QS signaling molecules 3-oxo-C12-HSL and C4-HSL; Attenuation of LasA, LasB, and LuxR.[129,130,131]
Methyl-gallate Suppression of the synthesis of the extracellular polymeric matrix, inhibition of QS signaling, and alteration of the microbial cell membrane.[132]
PhloretinApple, pear and strawberry fruitsDownregulation of virulence genes essential for surface attachment, biofilm formation, and QS.[133]
TerpenoidsCarvacrolOrganum vulgareReduction in QS signalling.[134]
Ursolic acidMany medical plantsInhibition of biofilm formation and disruption of cell membrane.[135]
4-epi-primaric acidAralia cachemirica L.Inhibition of biofilm formation and disruption of cell membrane.[136]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

D’Aquila, P.; De Rose, E.; Sena, G.; Scorza, A.; Cretella, B.; Passarino, G.; Bellizzi, D. Quorum Quenching Approaches against Bacterial-Biofilm-Induced Antibiotic Resistance. Antibiotics 2024, 13, 619. https://doi.org/10.3390/antibiotics13070619

AMA Style

D’Aquila P, De Rose E, Sena G, Scorza A, Cretella B, Passarino G, Bellizzi D. Quorum Quenching Approaches against Bacterial-Biofilm-Induced Antibiotic Resistance. Antibiotics. 2024; 13(7):619. https://doi.org/10.3390/antibiotics13070619

Chicago/Turabian Style

D’Aquila, Patrizia, Elisabetta De Rose, Giada Sena, Angelo Scorza, Bonaventura Cretella, Giuseppe Passarino, and Dina Bellizzi. 2024. "Quorum Quenching Approaches against Bacterial-Biofilm-Induced Antibiotic Resistance" Antibiotics 13, no. 7: 619. https://doi.org/10.3390/antibiotics13070619

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

Article Metrics

Back to TopTop