Next Article in Journal
Near-Surface Reconfiguration of Biopolymer Blends by Mechanical Embossment: Creation of Friction-Reduced Foils
Previous Article in Journal
Synthesis and Characterization of Bio-Based Polyurea Derived from Liquefied Wood of Wooden Musical Instrument Offcuts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Targeting Bacterial Communication: Evaluating Phytochemicals as LuxS Inhibitors to Disrupt Quorum Sensing

by
Yiannis Sarigiannis
1 and
Christos Papaneophytou
2,*
1
Department of Health Sciences, School of Life and Health Sciences, University of Nicosia, 2417 Nicosia, Cyprus
2
Department of Life Sciences, School of Life and Health Sciences, University of Nicosia, 2417 Nicosia, Cyprus
*
Author to whom correspondence should be addressed.
Macromol 2024, 4(4), 753-771; https://doi.org/10.3390/macromol4040045
Submission received: 1 September 2024 / Revised: 12 October 2024 / Accepted: 4 November 2024 / Published: 5 November 2024

Abstract

Bacterial quorum sensing (QS) is a critical communication process that regulates gene expression in response to population density, influencing activities such as biofilm formation, virulence, and antibiotic resistance. This study investigates the inhibitory effects of five phytochemicals—apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid—on the S-ribosylhomocysteinase (LuxS) enzyme, a key player in AI-2 signaling across both Gram-positive and Gram-negative bacteria. Using molecular docking studies, we identified that these phytochemicals interact with the LuxS enzyme, with apigenin, carnosol, chlorogenic acid, and rosmarinic acid binding within the substrate-binding pocket and exhibiting binding scores below −7.0 kcal/mol. Subsequent in vitro assays demonstrated that these compounds inhibited AI-2 signaling and biofilm formation in Escherichia coli MG1655 in a concentration-dependent manner. Notably, carnosol and chlorogenic acid showed the most potent effects, with IC50 values of approximately 60 μM. These findings suggest that these phytochemicals may serve as potential QS inhibitors, providing a foundation for developing new anti-pathogenic agents to combat bacterial infections without promoting antibiotic resistance. Further studies are warranted to explore the therapeutic applications of these compounds in both clinical and agricultural settings.

Graphical Abstract

1. Introduction

Quorum sensing (QS) is a bacterial communication mechanism that involves the production, detection, and response to extracellular signaling molecules known as auto-inducers (AIs) [1]. This process enables bacterial populations to coordinate behavior changes synchronously in response to shifts in density and species composition within their community. QS orchestrates gene expression programs that underpin collective behaviors, such as the expression of virulence factors in pathogenic bacteria [2]. By modulating gene expression based on cell density, bacteria can efficiently undertake energy-intensive activities, maximizing their impact on the environment or host. This collective capability ensures that energetically costly processes are initiated only when they are most likely to be effective [3]. Both Gram-positive (G+) and Gram-negative (G) bacteria utilize QS to regulate gene expression and coordinate collective behaviors. However, they employ different mechanisms and signaling molecules for this process [4].
G+ bacteria use small, post-translationally modified peptides called autoinducing peptides (AIPs) as signaling molecules [5]. These AIPs are often part of a histidine kinase two-component signal transduction system. In this system, secreted precursor AIPs are processed into mature AIPs by extracellular proteases [6]. These mature AIPs can be imported back into the cell and identified by cytoplasmic transcription factors. Typically, AIPs bind to membrane-bound sensor kinase receptors, which autophosphorylate and pass the phosphate to a cognate cytoplasmic response regulator. The phosphorylated response regulator then activates the transcription of genes in the QS regulon. Additionally, G+ bacteria secrete various peptides, including pheromones, that regulate specific genes involved in processes such as virulence factor biosynthesis, bacterial competence, and bacterial conjugation [7,8].
In contrast, G bacteria use small molecules as autoinducers (AIs), primarily acyl-homoserine lactones (AHLs) or other molecules synthesized from S-adenosylmethionine (SAM) [3]. AIs are produced within the cell and diffuse freely across bacterial membranes. When AI concentration is high, typically at a high cell density (HCD), they bind to specific receptors located either in the inner membrane or cytoplasm. These receptors are often transcription factors that regulate the expression of genes in the QS regulon [9]. In some cases, AIs are detected by two-component histidine kinase receptors similar to those in G+ bacteria [10]. A common feature in G quorum sensing is autoinduction, where the AI-driven activation of QS leads to increased AI synthesis, creating a feed-forward loop that promotes synchronous gene expression across the population [3].
The only QS mechanism that is employed by both G+ and G bacteria is the production of autoinducer-2 (AI-2) signaling molecules by the enzyme S-ribosylhomocysteinase (LuxS) (EC 4.4.1.21), which catalyzes the cleavage of S-ribosylhomocysteine (SRH) into homocysteine and 4,5-dihydroxy-2,3-pentanedione (DPD) [11]. This system enables collective decision making among bacterial populations regarding gene expression through the secretion and detection of AI-2 [12]. In Vibrio harveyi, AI-2 is identified as a furanosyl-borate-diester, whereas in Escherichia coli, it appears as a furan molecule [13]. The synthesis of AI-2 involves converting S-adenosylhomocysteine (SAH) into homocysteine through either a single-step reaction using SAH hydrolase (SahH) or a two-step process involving SAH nucleosidase (Pfs) and LuxS, which specifically cleaves the thioether linkage of SRH to produce 4,5-dihydroxy-2,3-pentanedione (DPD). DPD can then rearrange to form R- or S-2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (R- or S-THMF), collectively known as AI-2 [14,15]. The ability of different bacterial species to sense the AI-2 produced by others suggests its role as a potential universal signal for interspecies communication [16].
Furthermore, LuxS is prevalent in a diverse array of both G+ and G and is integral not only to the production of AI-2 but also to critical metabolic functions. It catalyzes the hydrolysis of S-adenosyl homocysteine into S-adenosylmethionine (SAMe), a key methyl donor. This activity supports the methyl cycle, vital for the synthesis of vitamins and polyamines within bacterial central metabolism [17]. The presence of the LuxS/AI-2 system underscores its importance as a global regulatory network, responding to bacterial population changes and significantly influencing cellular behavior [18]. Given the pivotal role of LuxS in this universal signaling system, targeting it with specific inhibitors can profoundly impact bacterial communication and pathogenicity. Thus, understanding the structural and functional nuances of LuxS is paramount given its universal role in QS across diverse bacterial species, leading directly into our investigations of specific inhibitors [19].
The crystal structures of LuxS homologues [20,21,22] reveal a homodimer architecture that is distinctive to a specific group of enzymes with related functions. These enzymes, including LuxS, typically engage in metabolic processes related to the handling of SAMe-derived molecules, playing roles in various bacterial signaling and metabolic pathways. This structural setup is particularly suited to their function in cleaving thioether bonds, a critical step in producing signaling molecules like AI-2, which is fundamental for bacterial communication. Each enzyme forms two identical active sites at the interface of the dimer. These sites incorporate residues from both subunits within a conserved HXXEH motif, where an Fe2+ ion is tetrahedrally coordinated. The enzyme’s catalytic activity, involving a series of proton-transfer reactions facilitated by the Fe2+ ion and two residues from LuxS, leads to the predicted chemical transformations [22]. These structural and functional characteristics confirm that LuxS is the enzyme previously identified as the ribosylhomocysteine-cleavage enzyme [22,23], which is known to catalyze the formation of DPD in Escherichia coli [24]. Examples of LuxS include but are not limited to that of Haemophilus influenzae (UniProt: P44007), Salmonella typhi (UniProt: Q8Z4D7), Vibrio cholerae (UniProt: Q9KUG4), Campylobacter concisus (UniProt: A7ZGE2), and Neisseria meningitidis (UniProt: A9M0P7).
QS is a pivotal mechanism in bacteria that regulates activities such as biofilm formation and virulence factor production. Notably, recent research indicates that QS systems may play a role in this resistance [25,26]. Globally, the rise of antibiotic-resistant bacterial pathogens poses a significant health challenge [27]. Targeting QS offers a novel antibacterial approach that not only prevents resistance development but also suppresses virulence factors dependent on bacterial population density [28]. By disrupting QS, this strategy aims to diminish bacterial virulence and disease-causing potential without promoting the emergence of resistance, as it does not exert selective growth pressure on bacteria [29]. Furthermore, combining anti-QS strategies with traditional antibiotics presents a promising avenue to tackle bacterial infections and the burgeoning problem of antibiotic resistance [30]. Anti-QS strategies encompass a wide array of approaches, including natural compounds, antibody-mediated quorum quenching (QQ), computer-aided drug design for QQ, repurposing FDA-approved drugs as anti-QS agents, and the modulation of quorum-sensing molecules [30]. Notably, phytochemicals from aromatic and medicinal plants have recently been recognized for their potential as QS inhibitors, with several compounds already identified as effective in disrupting QS [31,32].
We have recently explored the anti-QS properties of ethanolic extracts from eight aromatic plants native to Cyprus: Origanum vulgare subsp. hirtum (oregano), Rosmarinus officinalis (rosemary), Salvia officinalis (common sage), Lavandula spp. (lavender), Calendula officinalis (calendula), Melissa officinalis (lemon balm), Sideritis cypria (Cypriot mountain tea), and Aloysia citriodora (lemon verbena) [33]. Our findings revealed that the extracts from oregano, rosemary, and common sage exhibited the most potent inhibitory effects on AI-2 signaling. These extracts also significantly reduced biofilm formation by more than 60% and inhibited both swimming and swarming motilities of E. coli MG1655 in a concentration-dependent manner, distinguishing them as effective anti-QS agents. These activities were achieved without affecting bacterial growth, suggesting a mechanism of action that targets QS pathways rather than bacterial viability. The remaining plant extracts demonstrated varying degrees of inhibitory activity. Furthermore, we identified five key phytochemicals that were the major components in the ethanolic extracts that could be responsible for these effects: carnosol, chlorogenic acid, and quercetin were common to all three plants, while apigenin and rosmarinic acid were uniquely detected in the oregano extract. Given the critical role of the LuxS/AI-2 system in QS, we extend our investigation herein to determine the specific anti-QS activities of each identified phytochemical and their potential for selectively inhibiting the LuxS protein. By elucidating the specific interactions between these phytochemicals and the LuxS enzyme, we anticipate not only advancing our understanding of QS modulation but also paving the way for novel anti-QS therapies that are less likely to contribute to resistance. This ongoing research not only contributes to our understanding of plant-based QS inhibitors but also enhances the potential for developing novel therapeutic agents to combat infections with QS-dependent pathogens, particularly those resistant to conventional antibiotics.

2. Materials and Methods

2.1. LuxS Structure Prediction

Since the experimental (crystallographic) 3D structure of the LuxS protein from E. coli MG1655 (UniProt ID: P45578) is not available in the Protein Data Bank (PDB), we utilized the AlphaFold Protein Structure Database (https://alphafold.ebi.ac.uk/entry/P45578, accessed on 1 August 2024) to obtain the predicted structure of the protein [34]. AlphaFold provides a per-residue model confidence score (pLDDT) ranging from 0 to 100, which indicates the confidence level of each residue’s positional accuracy. The model confidence for E. coli LuxS is very high, as reflected by the pLDDT (>90) score. However, the predicted protein models in the AlphaFold database lack coordinates for small molecules essential for molecular structure or function. Specifically, in the case of LuxS, the Fe2+ ion is missing. To address this, we utilized AlphaFill (https://alphafill.eu/; accessed on 1 August 2024) [35] to add the missing ion using a known PDB complex template (PDB ID: 5E68) and optimizing it with the YASARA feature (Supplementary Figure S1). As the template protein features a Zn2⁺ ion at its active site, the model predicted by AlphaFill also initially featured the same ion. However, since E. coli LuxS contains Fe2+ as its metal ion, we addressed this issue by replacing the Zn2⁺ with a Fe2⁺ ion using ChimeraX software (UCSF ChimeraX v1.18, Resource for Biocomputing, Visualization, and Informatics, University of California, San Francisco, CA, USA).
Subsequently, we used the HDock server (http://hdock.phys.hust.edu.cn/, accessed on 1 August 2024) [36] to build the protein dimer. Among the generated models, “Model 0” was selected for further analysis (Supplementary Figure S2). This model was constructed using a known PDB complex template (PDB ID: 5E68), which HDock indicated as having high confidence. Due to its foundation on a validated PDB structure and the absence of the docking and confidence scores typically calculated for de novo models, “Model 0” was deemed the most reliable and was therefore utilized in all subsequent docking studies. The resulting protein model was visualized and analyzed using UCSF ChimeraX v1.18 (Resource for Biocomputing, Visualization, and Informatics, University of California, San Francisco, CA, USA) [37,38] to evaluate the structural integrity and potential functional sites.

2.2. Docking Studies

The predicted 3D model of LuxS was used to assess the binding mode of five phytochemicals: apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid. As mentioned above, these compounds were identified as the major components in the ethanolic extracts of medicinal plants from Cyprus flora [33]. Docking studies were performed using PyRx v1.1 [39], employing the AutoDock Vina option with an exhaustiveness level of 8. The protein was prepared for docking by adding polar hydrogens and correcting charges under physiological pH conditions (pH 7.4), as previously described [40], using the DockPrep feature of ChimeraX software.
The two-dimensional chemical structures of the phytochemicals in Structured Data Format (SDF) were retrieved from the PubChem-NCBI database (https://pubchem.ncbi.nlm.nih.gov/, accessed on 2 August 2024) [41]. The SDF files were minimized and converted into the PDB format using the OpenBabel feature of PyRx. The phytochemicals were then screened for blind docking across the entire surface of the protein dimer. The grid box was centered at x = 12, y = 1, and z = 12, with dimensions of 65 Å × 50 Å × 65 Å. The docked structures were retrieved and visualized using BIOVIA Discovery Studio 2021 (Dassault Systèmes BIOVIA, San Diego, CA, USA). All compounds exhibited a binding score of less than −6.5 kcal/mol and were evaluated further in vitro for their anti-QS potential, as detailed in the subsequent sections.

2.3. Chemicals, Bacterial Strains, Media, and Culture Conditions

Unless otherwise stated, all chemicals, including pure phytochemicals, were purchased from Merck (Darmstadt, Germany). These included apigenin (Cat. #10798), carnosol (Cat. #49702), chlorogenic acid (Cat. #C3878), quercetin (Cat. #Q4951), and rosmarinic acid (Cat. #R4033), all of analytical grade. Phytochemical stock solutions (10 mM) were prepared in DMSO.
E. coli MG 1655 (ATCC-700926) and V. harveyi BB-170 (ATCC-BAA-1117) were obtained from the American Type Culture Collection (ATCC; Wesel, Germany).
E. coli was grown in a Luria Bertani (LB) medium consisting of 1% tryptone, 0.5% yeast extract, and 1% NaCl at 37 °C. V. harveyi BB-170 was grown at 30 °C in an autoinducer bioassay (AB) medium (ATTC medium: 2034) consisting of (per 1 L) 17.53 g of ΝaCl, 6.02 g of MgCl2, and 2.0 g of casamino acids (vitamin-free). The pH of the medium was adjusted to pH 7.0 with 1 M KOH and autoclaved at 121 °C for 15 min. The solution was cooled to room temperature, and 10 mL of a 1 M potassium phosphate buffer (pH 7.0), 10 mL of a 0.1 M sterile arginine solution, and 20 mL of 50% v/v sterile glycerol were added to the medium.

2.4. Autoinducer-2 Bioassay

The AI-2 bioassay was performed as previously described [42,43], with minor modifications to assess the effects of five phytochemicals on AI-2 production in E. coli. In brief, E. coli was grown overnight at 37 °C in LB medium supplemented with 0.5% glucose. The following day, the overnight culture was used to inoculate (1:100) fresh LB medium containing 0.5% glucose. The culture was then incubated at 37 °C for 5 h with continuous shaking at 250 rpm [33]. To prepare the cell-free supernatant (CFS), the culture was centrifuged at 16,000× g for 15 min at 4 °C. The resulting supernatant was filtered through 0.2 μm syringe filters (VWR, West Chester, PA, USA), aliquoted, and stored at −20 °C until the AI-2 bioluminescence assay was conducted.
To assess the effects of the five phytochemicals on AI-2 production, the reporter strain V. harveyi BB170 was grown for 16 h at 30 °C in AB medium and subsequently diluted (1:5000) into fresh AB medium. A total of 90 μL of the diluted cells was added to the wells of a 96-well plate and mixed with 9 μL of CFS from E. coli MG1655 and 1 μL of each phytochemical solution at various concentrations (ranging from 5 to 200 μΜ), as described in the results section. Blank controls (9 μL of CFS + 1 μL of DMSO) and a negative control (9 μL of AB medium + 1 μL of DMSO) were included in each experiment.
The plates were incubated at 30 °C with continuous shaking at 100 rpm, and luminescence readings (in relative light units, RLU) were recorded every 20 min using a PerkinElmer VictorX3 2030 Multiplate Reader (PerkinElmer, Waltham, MA, USA) in chemiluminescence mode. The inhibition of AI-2 activity was expressed as a percentage relative to the blank control and calculated using the following equation (Equation (1)) [44].
% AI-2   inhibition = 1 RLU   of   sample RLU   of   blank   control × 100
The IC50 values (i.e., the concentration of inhibitor required to reduce the measured activity to 50% of its maximum value) for the tested phytochemicals were calculated using Prism v8.0.2 (GraphPad, La Jolla, CA, USA) by utilizing the software’s built-in equations and the raw data obtained from the enzymatic assays. These values were determined by fitting the dose–response curves with non-linear regression analysis.

2.5. Biofilm Formation Assay

The effects of each of the five phytochemicals on biofilm formation were assessed using sterile 96-well flat-bottom polystyrene plates, as previously described [33,45]. Positive controls (bacterial cells + LB), medium controls (LB only), and solvent controls (cells + LB + DMSO) were included in the experiments. All experiments were conducted in triplicate.
The appropriate concentration of each phytochemical was added to the test wells before inoculation. Plates were then incubated at 37 °C with continuous shaking at 100 rpm. Following 48 h of cultivation, the content of each well was discarded, and the wells were rinsed three times with phosphate-buffered saline (PBS). The wells were then fixed by drying for 1 h at 37 °C in the incubator. Once fully dry, 200 μL of 0.1% crystal violet stain was added to each well and incubated for 15 min at 25 °C. Excess dye was rinsed off with tap water. Subsequently 200 μL of 96% ethanol was added to each well to solubilize the stain adhering to the biofilm biomass. The ethanol containing the solubilized stain was transferred to clean wells of a 96-well plate, and the absorbance at 570 nm (A570 nm) was measured using a PerkinElmer VictorX3 2030 Multiplate Reader (PerkinElmer, Waltham, MA, USA). The biofilm inhibition rate was calculated using the following equation (Equation (2)).
% Biofilm   inhibition = 1 A 570   nm   of   the   sample A 570   nm   of   the   positive   control × 100

2.6. Effects of the Five Phytochemicals on Bacterial Growth

The effects of the five phytochemicals at the highest concentration tested (200 μΜ) on the growth of E. coli MG1655 were evaluated in liquid culture (200 μL) in the wells of a 96-well plate following 20 h of cultivation. Growth controls (bacteria cells + LB), medium controls (LB only), and solvent controls (cells + LB + DMSO) were used. All experiments were carried out in triplicate. Optical density values at 600 nm (OD600nm) were obtained using a Perkin Elmer VictorX3 2030 Multiplate reader (PerkinElmer, Waltham, MA, USA) at 0 and 20 h post-inoculation. The % growth on inhibition was determined using the following equation (Equation (3)) [46]:
%   inhibition = ( 1   OD t 20 h OD t 0 h OD gc 20 h OD gc 0 h )   ×   100
where ODt20h and ODt0h are the OD600nm values of the test well at 20 h and 0 h post-inoculation, respectively, while ODgc20h and ODgc0h are the OD600nm values of the growth control well at 20 h and 0 h post-inoculation, respectively.
The effects of five phytochemicals on bacterial growth were further assessed by performing viable plate counts [47]. Phytochemicals at a final concentration of 200 μΜ and cultures of E. coli (107 CFU/mL) were added to the wells of a 96-well plate (200 µL per well). Bacteria (107 CFU/mL) without any of the compounds were used as a control. The plates were incubated at 37 °C without shaking for 24 h. Following incubation, bacterial suspensions were transferred to clean Eppendorf tubes, centrifuged at 5000× g for 5 min at 4 °C, washed three times with PBS, and resuspended in 200 μL of fresh LB medium. Each suspension was subsequently serially diluted in LB and plated on LB agar. After incubation at 37 °C for 16 h, the number of viable bacteria was determined and expressed as CFU/mL.

2.7. Statistical Analysis

Unless otherwise stated, experiments were carried out in triplicate, and the data are presented as the mean values ± standard deviation (SD). A one-way ANOVA followed by Tukey’s multiple-comparison test was used to compare the effects of the five phytochemicals on the biofilm formation by E. coli MG1655. Statistical significance was set at p < 0.05. The statistical analysis was performed using GraphPad Prism (v.8.2, GraphPad Software Inc., San Diego, CA, USA).

3. Results

3.1. Predicted 3D Structure of LuxS

We have previously used E. coli MG1655 as a model organism to investigate the QS activities of ethanolic extracts from several aromatic plants, including oregano, rosemary, common sage, lavender, calendula, lemon balm, Cyprus mountain tea, and lemon beebrush. Notably, extracts from oregano, rosemary, and common sage were found to inhibit AI-2 signaling by over 60%. To further explore the impact of the major phytochemical components of these extracts on QS mechanisms and specifically on the LuxS/AI-2 signaling system, we obtained the predicted 3D structure of the LuxS protein in E. coli (Figure 1A). As aforementioned, because the experimental structure of the E. coli LuxS protein was not available in the PDB, we obtained its predicted 3D structure from AlphaFold (https://alphafold.ebi.ac.uk/entry/P45578; accessed on 1 August 2024). The missing ion (Fe2+) was added using AlphaFill while the LuxS dimer was developed with the HDOCK server, as described in Section 2.1. “LuxS Structure Prediction”. LuxS is a small metalloenzyme, consisting of 171 amino acids in E. coli, that forms homodimers with two identical active sites at the interface between subunits [48]. High-resolution crystal structures of LuxS, both in its free form and when bound to S-ribosyl-homocysteine (SRH), have been determined at 1.2 Å for various bacterial species [20,21,22]. These studies confirm the enzyme’s consistent homodimeric structure and reveal each active site’s coordination of a divalent metal ion. Initially thought to contain Zn2+ based on inductively coupled plasma (ICP) metal analysis, further research by Pei and colleagues identified Fe2+ as the actual native metal [49]. The active sites are highly conserved across different species, characterized by a similar overall fold and an invariant HXXEH motif, which is typically found in Zn2+-binding proteins. The metal ion is tetrahedrally coordinated by the side chains of His 54 and His 58 in the HXXEH motif, a cysteine residue (Cys128 in E. coli; Supplementary Figure S1), and a water molecule.
Several conserved residues from the symmetry-related chain contribute to the lining of the cavity that likely accommodates the substrate (Figure 1B). Notably, both chains of the enzyme contribute amino acids to form the binding site. The “A” chain provides catalytic residues, including His54, Glu57, and His58 from the HXXEH motif, as well as Cys128. Additionally, four conserved residues from the “B” subunit—Ser6, Phe7, His11, and Arg39—are located in the active site and are potentially important for substrate binding and/or catalysis [24]. It is important to note that Glu57 and Cys83 (Cys-84 in Bacillus subtilis LuxS) are crucial residues for LuxS catalysis, as indicated by mutagenesis studies [49].
The sequence alignment (Figure 1C) of E. coli LuxS with LuxS proteins from other bacterial species highlights the conserved residues involved in binding and catalysis. In addition to the aforementioned conserved residues, other amino acids that have been implicated in binding and/or catalysis, including Phe80 [48], Gly82, and Gly93 [50], are also noted.

3.2. Molecular Docking

We subsequently employed molecular docking studies to explore the interaction and binding modes of the five phytochemicals—apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid—that were identified as the major components in the ethanolic extracts of medicinal plants from the Cyprus flora, as previously described [33], with the predicted LuxS dimer. The structures of the five tested compounds are illustrated in Figure 2.
By performing blind docking across the entire surface of the protein, all the tested compounds except quercetin were found to bind within the same active pocket in the cavity between the two chains (Figure 3). We further reviewed the docking poses generated for quercetin to ascertain if any structures were located in the same active site as the other compounds. Notably, our analysis confirmed that none of the quercetin poses were located in the active site or interacted with the residues crucial for catalysis, suggesting a distinct mode of interaction. The binding scores of the five compounds ranged from −6.7 (quercetin) to −8.3 kcal/mol (rosmarinic acid), as illustrated in Table 1.
The molecular interactions of the five phytochemicals with the amino acids in the active pocket of LuxS are illustrated as 2D diagrams in Figure 4. Most amino acids in the active pocket formed hydrogen bonds and van der Waals interactions with the selected natural products. Additionally, other types of interactions, such as pi-anion and pi-sigma bonds, were predicted, as shown in Figure 4.
Consistent with previous studies, both chains of the LuxS dimer were involved in substrate (phytochemical) binding. Chain A contributed His54, Glu57, His58, Glu122, and Cys128, although not all these residues participated in the binding of each phytochemical, as shown in Table 1 and Figure 4. Similarly, Chain B was involved with Ser6, Phe7, His11, Arg39, and Cys83, but not all these conserved amino acids were implicated in binding every substrate (Table 1 and Figure 4).
However, the key conserved residues previously reported to be involved in substrate binding—Phe80 [48] Gly82, and Gly93 [50]—did not make significant contributions in binding the tested phytochemicals, possibly due to the use of a different strain. Notably, His11 in Chain B participated in the binding of all phytochemicals except quercetin, underscoring the importance of this residue in substrate binding. A mutation of His11 has been shown to result in a ~1000-fold reduction in LuxS activity [51]. Interestingly, none of the conserved amino acids were implicated in the binding of quercetin (binding score of −6.7 kcal/mol), as it had been placed on a different area on the dimer.

3.3. Effects of Phytochemicals on AI-2 Signaling

We subsequently evaluated the anti-QS potential of the five phytochemicals by monitoring AI-2 inhibition using the well-established V. harveyi BB170 bioassay [52]. This assay relies on the ability of the reporter strain V. harveyi BB170 to produce bioluminescence specifically in response to AI-2. At cell densities of 106 to 107 CFU/mL, bioluminescence can be detected when AI-2 is added (spiked) [53].
In our experiments, V. harveyi BB170 was incubated with a known concentration of exogenous AI-2 (i.e., cell-free supernatant (CFS) from an E. coli culture) to induce luminescence, along with either a phytochemical or its respective blank medium (CFS without the phytochemical). Inhibition was determined to have occurred when the luminescence of the sample was lower than that of its corresponding blank. Based on literature precedents, phytochemicals are typically tested at concentrations ranging from 5 to 200 μM [53,54,55]; therefore, we assessed the effects of the five phytochemicals on AI-2 production within this range.
Previous studies have shown that the concentration of the AI-2 signaling molecule in E. coli peaks during the mid-to-late growth phase, followed by a significant decrease during the stationary phase [56]. In our earlier work [33], we also demonstrated that the concentration of AI-2 in E. coli MG1655 increased with incubation time up to 6 h but began to decrease afterward. Based on these observations, in this work, we used CFSs from E. coli cultures grown for 5 h at 37 °C in the presence of 0.5% glucose.
The results illustrated in Figure 5 demonstrate that apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid inhibit AI-2 signaling in a concentration-dependent manner with varying degrees of efficacy. Notably, the IC50 values for these compounds provide valuable insights into their relative potencies. Carnosol and chlorogenic acid exhibited similar IC50 values of 58.04 μM and 60.52 μM, respectively, indicating that they are among the more potent inhibitors of AI-2 signaling. Apigenin, in contrast, showed a higher IC50 value of 70.63 μM, likely due to its lack of interaction with key catalytic residues such as Ang39 and Cys128, as indicated by our docking studies, which are essential for LuxS activity.
Quercetin displayed an IC50 value greater than 200 μM, indicating a minimal inhibitory effect under the tested conditions. This lack of efficacy might be due to its poor interaction with key amino acid residues within the LuxS active site, as suggested by the docking results (Figure 3 and Figure 4). Interestingly, despite rosmarinic acid showing the strongest binding score in the docking studies (−8.3 kcal/mol), its IC50 value was higher (80.63 μM) than those of carnosol and chlorogenic acid. This suggests that while binding score (energy) is important, the specific interactions with crucial residues, such as Cys83, are also vital for effective inhibition. The docking studies further indicate that rosmarinic acid does not interact with Cys83, unlike carnosol and chlorogenic acid, which may explain its comparatively lower inhibitory potency despite its strong binding score.
These findings emphasize that both the binding energy and specific interactions with key residues are crucial in determining the effectiveness of these compounds in inhibiting AI-2 signaling. Compounds like carnosol and chlorogenic acid, which interact with critical residues such as Cys83 and Cys128, demonstrate more potent inhibitory effects, underscoring the importance of these interactions in the LuxS catalytic mechanism. Overall, the data suggest that while all tested compounds can bind to the LuxS enzyme, their ability to inhibit AI-2 signaling is significantly influenced by their interactions with specific key residues that are essential for LuxS activity.

3.4. Effects of Phytochemicals on Biofilm Formation

Previous studies have demonstrated that QS plays a crucial role in biofilm formation and differentiation in various bacterial species, including E. coli [57,58]. Notably, AI-2 signaling is closely associated with biofilm formation in many bacteria [59,60], including E. coli [61]. It has been shown that AI-2 production and uptake influence several E. coli phenotypes, such as biofilm formation, motility, and virulence [62,63]. During autoaggregation or biofilm formation, AI-2 functions as a chemoattractant, drawing planktonic cells to growing cell aggregates [61,64].
Given the significant role of AI-2 signaling in biofilm formation, we investigated the effects of five phytochemicals—apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid—on the formation of E. coli MG1655 biofilms. Biofilm formation was quantified using crystal violet staining at final concentrations of 50, 100, and 200 μM, as illustrated in Figure 6. The results indicate that all five compounds inhibited biofilm formation to varying degrees. Specifically, apigenin reduced biofilm formation by approximately 50% at 50 μM, with further reductions observed at 100 and 200 μM, though these differences were not statistically significant. Among the tested compounds, carnosol showed the highest inhibitory effect, reducing biofilm formation by over 60% even at the lowest concentration of 50 μM, with only a slight additional reduction at higher concentrations. Chlorogenic acid inhibited biofilm formation in a concentration-dependent manner. Although quercetin had only a minimal effect on AI-2 signaling (as shown in Figure 5D), it still significantly reduced biofilm formation, with increased inhibition at higher concentrations. Lastly, rosmarinic acid also displayed a concentration-dependent inhibition of biofilm formation.

3.5. Effects of the Five Phytochemicals on Bacterial Growth

The main difference between QS inhibitors (QSIs) and antibiotics is that QSIs inhibit the production of virulence factors without killing pathogenic bacteria, thereby reducing the likelihood of bacteria developing drug-resistant mutations [33]. We have previously demonstrated that the ethanolic extracts of oregano, rosemary, and common sage, which contain the five compounds tested here as major components, had minimal impact on bacterial growth [33]. Therefore, in this study, we further investigated whether the observed anti-QS activities of the five compounds (at the maximum concentration tested, viz. 200 μΜ) were correlated with bactericidal activity by examining their effects on the growth of E. coli MG1655.
Consistent with our previous observations [33], none of the phytochemicals exhibited any bactericidal activity, as determined by the inhibition-of-growth assay (Figure 7A) and viable plate counts (Figure 7B).

4. Discussion

Bacteria use QS to coordinate gene expression programs that underlie collective behaviors. QS is a cell–cell communication process that depends on the production, release, detection, and group-level response to extracellular signaling molecules known as AIs. These small chemical signals regulate a wide variety of cellular activities, including motility, virulence, antibiotic production, biosurfactant formation, and biofilm formation [3,65].
AIs vary among bacterial species. For example, acyl-homoserine lactones are typically produced by G bacteria, while oligopeptides are secreted by G+ bacteria. These molecules primarily facilitate intraspecies communication [66]. However, AI-2 is a well-conserved QS signal synthesized by a wide range of G+ and G bacteria. AI-2, produced via the activated methyl cycle by the enzymes Pfs and LuxS, enables communication at both intra- and interspecies levels, thereby promoting coordination among different bacterial species [67,68].
Inhibiting QS pathways among bacteria presents a potential strategy for combating multi-drug-resistant bacterial infections. Compounds that can inhibit QS activity, particularly without affecting bacterial growth rates, are considered potential QS inhibitors [69]. Notably, several plant extracts and essential oils have been reported to exhibit both antimicrobial and anti-QS activities, highlighting the importance of identifying anti-QS compounds from natural sources, such as aromatic plants [70].
Considering the central role of LuxS in AI-2 signaling, in this study, we examined the inhibitory potential of five phytochemicals—apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid—previously identified as the major components of ethanolic extracts from oregano, rosemary, and common sage in the flora of Cyprus. We initially employed molecular docking studies to explore the interaction and binding models between the five phytochemicals and LuxS. Except for quercetin, the other four compounds were positioned in the cavity between the two protein subunits (chains) and exhibited binding scores of lower than −7.0 kcal/mol. Consistent with previous studies, both subunits contributed amino acids for substrate binding, including His54, Glu57, His58, and Cys128 from Chain A and Ser6, Phe7, His11, Arg39, and Cys83 from Chain B (Table 1 and Figure 4). In contrast to Meng et al. [71], who tested various natural products as inhibitors of LuxS in Lactobacillus reuteri and reported metal complexes between some compounds (e.g., norathyriol, mangiferin, and kaempferol) and Zn2⁺, we did not detect any complexes between the Fe2⁺ ion of E. coli LuxS and the five phytochemicals. This discrepancy is likely due to differences in the docking methods used (Schrödinger software version 2021-1 in Meng’s study vs. AutoDock in our study), the bacterial strains tested (L. reuteri vs. E. coli), and the specific phytochemicals examined.
We subsequently assessed the effects of the five phytochemicals on AI-2 signaling in vitro using the well-established V. harveyi bioassay [52] with the reporter strain BB170 which is highly sensitive to AI-2 (with a QS phenotype of AI-1, AI-2+), allowing even low levels of AI-2 to be detected in this bioassay. Inhibition was considered significant when the luminescence of a tested compound (i.e., phytochemical) was lower than that of the corresponding blank control. Consistent with our docking studies, apigenin, carnosol, chlorogenic acid, and rosmarinic acid inhibited AI-2 signaling in a concentration-dependent manner (Figure 5). Carnosol and chlorogenic acid were the most potent, with IC50 values of approximately 60 μM, while apigenin and rosmarinic acid exhibited slightly higher IC50 values of around 70 μM and 80 μM, respectively, despite apigenin having the lowest binding score (−8.3 kcal/mol) in our docking studies. This discrepancy can be partly explained by the fact that carnosol and chlorogenic acid interact with both Cys83 of Chain B and Cys128 of Chain A, which have been identified as essential amino acid residues for LuxS activity. However, the exact mechanism by which each of these phytochemicals inhibits LuxS activity is currently under investigation in our laboratory.
The five phytochemicals demonstrated a dose-dependent inhibition of biofilm formation under the experimental conditions (Figure 6). Although quercetin did not significantly impact AI-2 signaling—likely because it binds to a site away from the active site according to our docking studies (Figure 3 and Figure 4) and does not interact with any of the essential amino acids involved in substrate binding—it still had a significant effect on biofilm formation in E. coli. Additionally, quercetin exhibited only a minor impact on LuxS activity (Figure 5). To ensure that the observed inhibition of biofilm formation was not due to reduced bacterial growth, we assessed the impact of each phytochemical on E. coli growth separately and found that none of them inhibited bacterial growth (Figure 7).
To the best of our knowledge, while the five phytochemicals examined in this study have been explored for their general antimicrobial and anti-QS activities, there is limited direct research on their specific effects on LuxS activity and AI-2 signaling. Most available studies have focused on broader quorum sensing and biofilm inhibition without directly measuring LuxS activity or AI-2 levels. It should be noted that that the actions of phytochemicals against QS are diverse, multiple, and depend on the bacterial strain tested and the experimental method used.
Flavonoids such as quercetin [72] and apigenin [73] have been shown to act as quorum-sensing (QS) inhibitors, suppressing bacterial cell-to-cell communication and exerting antagonistic effects on bacterial signaling. Both compounds have been demonstrated to inhibit biofilm formation in E. coli O157 and V. harveyi [74]. Furthermore, quercetin has been found to inhibit QS-controlled virulence factors such as violacein, elastase, and pyocyanin in Chromobacterium violaceum CV12472 and P. aeruginosa PAO1 [75]. It has also been found to reduce the expression levels of QS-related genes such as lasI, lasR, rhlI, and rhlR at a concentration of 16 μg/mL. Additionally, quercetin inhibits QS circuitry by interacting with the transcriptional regulator LasR in P. aeruginosa [76].
The anti-QS properties of rosmarinic acid and chlorogenic acid have also been documented [77]. For example, rosmarinic acid was shown to inhibit biofilm formation and reduce the QS-mediated production of hemolysin, lipase, and elastase in Aeromonas hydrophila strains. It also downregulated virulence genes such as ahh1, aerA, lip, and ahyB [78].
In P. aeruginosa, rosmarinic acid has been shown to induce QS-dependent gene expression, increase biofilm formation, and enhance the production of virulence factors like pyocyanin and elastase. Another study revealed that rosmarinic acid induced the expression of 128 genes, including numerous virulence factor genes, and triggered a broad QS response in Pseudomonas aeruginosa PAO1. It also induced seven small RNAs (sRNAs), all of which were encoded near QS-induced genes [79].
Chlorogenic acid has been shown to inhibit biofilm formation in P. aeruginosa, suppress swarming motility, and reduce the production of virulence factors, including protease, elastase, rhamnolipids, and pyocyanin [80]. It has also exhibited similar inhibitory effects in Chromobacterium violaceum, reducing biofilm formation, swarming motility, chitinolytic activity, and violacein production. Specifically, chlorogenic acid was found to suppress QS in Chromobacterium violaceum (CECT 494) by inhibiting violacein production [81].
Likewise, direct studies on carnosol’s effects on LuxS activity or AI-2 signaling are limited; however, it has been observed to influence quorum-sensing mechanisms more broadly. For example, carnosol has been demonstrated to inhibit the Staphylococcus aureus quorum-sensing agr operon at a low concentration of 5 μM [82].

5. Conclusions

Bacterial resistance and pathogenesis are closely linked to their ability to sense and respond to population density through QS. The inhibition of the QS system is considered a novel strategy for developing antipathogenic agents, particularly for combating drug-resistant bacterial infections. Several mechanisms can achieve QS inhibition using natural compounds (phytochemicals), including [83,84,85]: (i) the inhibition of the biosynthesis of QS signaling molecules; (ii) competitive inhibition, where certain natural compounds share structural similarities with QS signaling molecules and compete for binding to receptor proteins, thereby disrupting the signal transduction pathways; and (iii) the inhibition of QS signal reception. Notably, phytochemicals exhibit diverse and multiple actions against QS, with their effects varying depending on the bacterial strain and experimental conditions used. In this study, we demonstrated for the first time that four phytochemicals—apigenin, carnosol, chlorogenic acid, and rosmarinic acid—abundant in ethanolic extracts from aromatic plants bind to the substrate-binding pocket of the LuxS enzyme of E. coli MG1655, thereby inhibiting AI-2 signaling. Importantly, AI-2 is a signal molecule produced by LuxS, an enzyme found in many bacterial species, and is essential for enabling interspecies communication.
These findings highlight the potential of these phytochemicals as promising candidates for further research in vitro and in vivo, as they may offer significant potential for new drug development aimed at mitigating bacterial resistance and pathogenesis through QS inhibition. Expanding our understanding of how these compounds interact with bacterial QS systems could lead to the development of novel therapeutic agents capable of targeting resistant bacterial strains without the selective pressure that often leads to antibiotic resistance.
It should be noted that besides anti-QS activity, polyphenolic compounds including apigenin, carnosol, chlorogenic acid, quercetin, and rosmarinic acid are also well known for their strong antioxidant properties and extensive pharmacological profiles. Their ability to interact with multiple and diverse molecular targets, known as promiscuity, presents both advantages and challenges. On the one hand, this characteristic enables these molecules to exert a wide range of therapeutic effects, such as anti-inflammatory, anticancer, and cardiovascular protective activities. However, it also complicates the prediction of side effects and interactions with other drugs, as these compounds can bind to various enzymes, receptors, and other cellular components. This complexity can hinder the therapeutic application of polyphenols, as their beneficial properties may also lead to unintended off-target effects. For example, while quercetin’s modulation of several signaling pathways can be beneficial in cancer treatment, it may also inadvertently affect metabolic pathways. Therefore, a thorough understanding of the promiscuity of these compounds is essential to fully exploit their therapeutic potential while minimizing adverse effects. Thus, research on the potential uses of polyphenols as therapeutic agents must be vigilant, considering both the polyvalent nature of these compounds and their extensive implications in pharmacodynamics.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/macromol4040045/s1, Figure S1: AlphaFill model of the E. coli LuxS enzyme with a Fe2⁺ ion; Figure S2: HDOCK server results for the docking simulations of the E. coli LuxS dimer.

Author Contributions

Conceptualization, Y.S. and C.P.; methodology, Y.S. and C.P.; software, C.P.; validation, Y.S. and C.P.; formal analysis, Y.S. and C.P.; investigation, C.P.; resources, Y.S. and C.P.; data curation, Y.S. and C.P.; writing—original draft preparation, C.P.; writing—review and editing, Y.S. and C.P.; visualization, Y.S. and C.P.; supervision, C.P.; project administration, C.P.; funding acquisition, C.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Nicosia/UNIVERSITAS Foundation Fund: Internal Funding for Research Projects, grant number: SeedGrant/2020/07, entitled “Identification and validation of natural quorum-sensing inhibitors to replace antibiotics in animal farms” (Acronym: NatuFarm).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author [C.P.] on special request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mukherjee, S.; Bassler, B.L. Bacterial quorum sensing in complex and dynamically changing environments. Nat. Rev. Microbiol. 2019, 17, 371–382. [Google Scholar] [CrossRef] [PubMed]
  2. 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]
  3. Papenfort, K.; Bassler, B.L. Quorum sensing signal–response systems in Gram-negative bacteria. Nat. Rev. Microbiol. 2016, 14, 576–588. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, L.; Luo, Y. Bacterial quorum-sensing systems and their role in intestinal bacteria-host crosstalk. Front. Microbiol. 2021, 12, 611413. [Google Scholar] [CrossRef]
  5. Lyon, G.J.; Novick, R.P. Peptide signaling in Staphylococcus aureus and other Gram-positive bacteria. Peptides 2004, 25, 1389–1403. [Google Scholar] [CrossRef]
  6. Håvarstein, L.S.; Coomaraswamy, G.; Morrison, D.A. An unmodified heptadecapeptide pheromone induces competence for genetic transformation in Streptococcus pneumoniae. Proc. Natl. Acad. Sci. USA 1995, 92, 11140–11144. [Google Scholar] [CrossRef]
  7. Banerjee, G.; Ray, A.K. Quorum-sensing network-associated gene regulation in Gram-positive bacteria. Acta Microbiol. Immunol. Hung. 2017, 64, 439–453. [Google Scholar] [CrossRef]
  8. Prazdnova, E.V.; Gorovtsov, A.V.; Vasilchenko, N.G.; Kulikov, M.P.; Statsenko, V.N.; Bogdanova, A.A.; Refeld, A.G.; Brislavskiy, Y.A.; Chistyakov, V.A.; Chikindas, M.L. Quorum-sensing inhibition by Gram-positive bacteria. Microorganisms 2022, 10, 350. [Google Scholar] [CrossRef] [PubMed]
  9. Galloway, W.R.; Hodgkinson, J.T.; Bowden, S.D.; Welch, M.; Spring, D.R. Quorum sensing in Gram-negative bacteria: Small-molecule modulation of AHL and AI-2 quorum sensing pathways. Chem. Rev. 2011, 111, 28–67. [Google Scholar] [CrossRef]
  10. Giannakara, M.; Koumandou, V.L. Evolution of two-component quorum sensing systems. Access Microbiol. 2022, 4, 000303. [Google Scholar] [CrossRef]
  11. Schauder, S.; Bassler, B.L. The languages of bacteria. Genes. Dev. 2001, 15, 1468–1480. [Google Scholar] [CrossRef] [PubMed]
  12. Vendeville, A.; Winzer, K.; Heurlier, K.; Tang, C.M.; Hardie, K.R. Making ‘sense’ of metabolism: Autoinducer-2, LUXS and pathogenic bacteria. Nat. Rev. Microbiol. 2005, 3, 383–396. [Google Scholar] [CrossRef] [PubMed]
  13. Surette, M.G.; Miller, M.B.; Bassler, B.L. Quorum sensing in Escherichia coli, Salmonella typhimurium, and Vibrio harveyi: A new family of genes responsible for autoinducer production. Proc. Natl. Acad. Sci. USA 1999, 96, 1639–1644. [Google Scholar] [CrossRef]
  14. Hu, X.; Wang, Y.; Gao, L.; Jiang, W.; Lin, W.; Niu, C.; Yuan, K.; Ma, R.; Huang, Z. The Impairment of Methyl Metabolism From luxS Mutation of Streptococcus mutans. Front. Microbiol. 2018, 9, 404. [Google Scholar] [CrossRef]
  15. García-Contreras, R.; Nuñez-López, L.; Jasso-Chávez, R.; Kwan, B.W.; Belmont, J.A.; Rangel-Vega, A.; Maeda, T.; Wood, T.K. Quorum sensing enhancement of the stress response promotes resistance to quorum quenching and prevents social cheating. ISME J. 2015, 9, 115–125. [Google Scholar] [CrossRef]
  16. Thompson, J.A.; Oliveira, R.A.; Djukovic, A.; Ubeda, C.; Xavier, K.B. Manipulation of the quorum sensing signal AI-2 affects the antibiotic-treated gut microbiota. Cell Rep. 2015, 10, 1861–1871. [Google Scholar] [CrossRef] [PubMed]
  17. Challan Belval, S.; Gal, L.; Margiewes, S.; Garmyn, D.; Piveteau, P.; Guzzo, J. Assessment of the roles of LuxS, S-ribosyl homocysteine, and autoinducer 2 in cell attachment during biofilm formation by Listeria monocytogenes EGD-e. Appl. Environ. Microbiol. 2006, 72, 2644–2650. [Google Scholar] [CrossRef]
  18. De Keersmaecker, S.C.J.; Sonck, K.; Vanderleyden, J. Let LuxS speak up in AI-2 signaling. Trends Microbiol. 2006, 14, 114–119. [Google Scholar] [CrossRef]
  19. Hardie, K.R.; Heurlier, K. Establishing bacterial communities by ‘word of mouth’: LuxS and autoinducer 2 in biofilm development. Nat. Rev. Microbiol. 2008, 6, 635–643. [Google Scholar] [CrossRef]
  20. Hilgers, M.T.; Ludwig, M.L. Crystal structure of the quorum-sensing protein LuxS reveals a catalytic metal site. Proc. Natl. Acad. Sci. USA 2001, 98, 11169–11174. [Google Scholar] [CrossRef]
  21. Lewis, H.A.; Furlong, E.B.; Laubert, B.; Eroshkina, G.A.; Batiyenko, Y.; Adams, J.M.; Bergseid, M.G.; Marsh, C.D.; Peat, T.S.; Sanderson, W.E.; et al. A structural genomics approach to the study of quorum sensing: Crystal structures of three LuxS orthologs. Structure 2001, 9, 527–537. [Google Scholar] [CrossRef] [PubMed]
  22. Ruzheinikov, S.N.; Das, S.K.; Sedelnikova, S.E.; Hartley, A.; Foster, S.J.; Horsburgh, M.J.; Cox, A.G.; McCleod, C.W.; Mekhalfia, A.; Blackburn, G.M.; et al. The 1.2 A structure of a novel quorum-sensing protein, Bacillus subtilis LuxS. J. Mol. Biol. 2001, 313, 111–122. [Google Scholar] [CrossRef]
  23. Miller, C.H.; Duerre, J.A. S-ribosylhomocysteine cleavage enzyme from Escherichia coli. J. Biol. Chem. 1968, 243, 92–97. [Google Scholar] [CrossRef]
  24. Pei, D.; Zhu, J. Mechanism of action of S-ribosylhomocysteinase (LuxS). Curr. Opin. Chem. Biol. 2004, 8, 492–497. [Google Scholar] [CrossRef] [PubMed]
  25. Xu, G.M. Relationships between the regulatory systems of quorum sensing and multidrug resistance. Front. Microbiol. 2016, 7, 958. [Google Scholar] [CrossRef]
  26. Tavío, M.M.; Aquili, V.D.; Fàbrega, A.; Vila, J.; Poveda, J.B. Overexpression of the quorum-sensing regulator sdiA and soxS is involved in low-level multidrug resistance induced in Escherichia coli AG100 by haloperidol, diazepam and NaCl. Int. J. Antimicrob. Agents 2012, 39, 91–93. [Google Scholar] [CrossRef] [PubMed]
  27. Bidault, P.; Chandad, F.; Grenier, D. Risk of bacterial resistance associated with systemic antibiotic therapy in periodontology. J. Can. Dent. Assoc. 2007, 73, 721–725. [Google Scholar] [PubMed]
  28. Krašovec, R.; Belavkin, R.V.; Aston, J.A.; Channon, A.; Aston, E.; Rash, B.M.; Kadirvel, M.; Forbes, S.; Knight, C.G. Where antibiotic resistance mutations meet quorum-sensing. Microb. Cell 2014, 1, 250–252. [Google Scholar] [CrossRef]
  29. Rasko, D.A.; Sperandio, V. Anti-virulence strategies to combat bacteria-mediated disease. Nat. Rev. Drug Discov. 2010, 9, 117–128. [Google Scholar] [CrossRef]
  30. Elfaky, M.A. Unveiling the hidden language of bacteria: Anti-quorum sensing strategies for gram-negative bacteria infection control. Arch. Microbiol. 2024, 206, 124. [Google Scholar] [CrossRef]
  31. Zhong, L.; Ravichandran, V.; Zhang, N.; Wang, H.; Bian, X.; Zhang, Y.; Li, A. Attenuation of Pseudomonas aeruginosa quorum sensing by natural products: Virtual screening, evaluation and biomolecular interactions. Int. J. Mol. Sci. 2020, 21, 2190. [Google Scholar] [CrossRef] [PubMed]
  32. Chakraborty, P.; Dastidar, D.G.; Paul, P.; Dutta, S.; Basu, D.; Sharma, S.R.; Basu, S.; Sarker, R.K.; Sen, A.; Sarkar, A.; et al. Inhibition of biofilm formation of Pseudomonas aeruginosa by caffeine: A potential approach for sustainable management of biofilm. Arch. Microbiol. 2020, 202, 623–635. [Google Scholar] [CrossRef] [PubMed]
  33. Panayi, T.; Sarigiannis, Y.; Mourelatou, E.; Hapeshis, E.; Papaneophytou, C. Anti-quorum-sensing potential of ethanolic extracts of aromatic plants from the flora of Cyprus. Plants 2022, 11, 2632. [Google Scholar] [CrossRef] [PubMed]
  34. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  35. Hekkelman, M.L.; de Vries, I.; Joosten, R.P.; Perrakis, A. AlphaFill: Enriching AlphaFold models with ligands and cofactors. Nat. Methods 2023, 20, 205–213. [Google Scholar] [CrossRef]
  36. Yan, Y.; Tao, H.; He, J.; Huang, S.-Y. The HDOCK server for integrated protein–protein docking. Nat. Protoc. 2020, 15, 1829–1852. [Google Scholar] [CrossRef]
  37. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera-a visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef]
  38. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Meng, E.C.; Couch, G.S.; Croll, T.I.; Morris, J.H.; Ferrin, T.E. UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 2021, 30, 70–82. [Google Scholar] [CrossRef]
  39. Dallakyan, S.; Olson, A.J. Small-molecule library screening by docking with PyRx. In Chemical Biology: Methods and Protocols; Hempel, J.E., Williams, C.H., Hong, C.C., Eds.; Springer: New York, NY, USA, 2015; pp. 243–250. [Google Scholar]
  40. Theerawatanasirikul, S.; Lekcharoensuk, P. Virtual screening of natural compounds targeting proteases of coronaviruses and picornaviruses. In In Silico Modeling of Drugs Against Coronaviruses: Computational Tools and Protocols; Roy, K., Ed.; Springer: New York, NY, USA, 2021; pp. 661–681. [Google Scholar]
  41. Sayers, E.W.; Barrett, T.; Benson, D.A.; Bolton, E.; Bryant, S.H.; Canese, K.; Chetvernin, V.; Church, D.M.; DiCuccio, M.; Federhen, S.; et al. Database resources of the National Center for Biotechnology Information. Nucleic Acids Res. 2011, 39, D38–D51. [Google Scholar] [CrossRef]
  42. Surette, M.G.; Bassler, B.L. Regulation of autoinducer production in Salmonella typhimurium. Mol. Microbiol. 1999, 31, 585–595. [Google Scholar] [CrossRef]
  43. Surette, M.G.; Bassler, B.L. Quorum sensing in Escherichia coli and Salmonella typhimurium. Proc. Natl. Acad. Sci. USA 1998, 95, 7046–7050. [Google Scholar] [CrossRef] [PubMed]
  44. Girennavar, B.; Cepeda, M.L.; Soni, K.A.; Vikram, A.; Jesudhasan, P.; Jayaprakasha, G.K.; Pillai, S.D.; Patil, B.S. Grapefruit juice and its furocoumarins inhibits autoinducer signaling and biofilm formation in bacteria. Int. J. Food Microbiol. 2008, 125, 204–208. [Google Scholar] [CrossRef] [PubMed]
  45. Merritt, J.H.; Kadouri, D.E.; O’Toole, G.A. Growing and analyzing static biofilms. Curr. Protoc. Microbiol. 2006, 22, 1B.1.1–1B.1.17. [Google Scholar] [CrossRef] [PubMed]
  46. Quave, C.L.; Plano, L.R.W.; Pantuso, T.; Bennett, B.C. Effects of extracts from Italian medicinal plants on planktonic growth, biofilm formation and adherence of methicillin-resistant Staphylococcus aureus. J. Ethnopharmacol. 2008, 118, 418–428. [Google Scholar] [CrossRef] [PubMed]
  47. Manner, S.; Fallarero, A. Screening of natural product derivatives identifies two structurally related flavonoids as potent quorum sensing inhibitors against gram-negative bacteria. Int. J. Mol. Sci. 2018, 19, 1346. [Google Scholar] [CrossRef]
  48. Wang, Y.; Yi, L.; Wang, S.; Fan, H.; Ding, C.; Mao, X.; Lu, C. Crystal structure and identification of two key amino acids Involved in AI-2 production and biofilm formation in Streptococcus suis LuxS. PLoS ONE 2015, 10, e0138826. [Google Scholar] [CrossRef]
  49. Zhu, J.; Dizin, E.; Hu, X.; Wavreille, A.S.; Park, J.; Pei, D. S-Ribosylhomocysteinase (LuxS) is a mononuclear iron protein. Biochemistry 2003, 42, 4717–4726. [Google Scholar] [CrossRef]
  50. Plummer, P.; Zhu, J.; Akiba, M.; Pei, D.; Zhang, Q. Identification of a key amino acid of LuxS involved in AI-2 production in Campylobacter jejuni. PLoS ONE 2011, 6, e15876. [Google Scholar] [CrossRef]
  51. Gopishetty, B.; Zhu, J.; Rajan, R.; Sobczak, A.J.; Wnuk, S.F.; Bell, C.E.; Pei, D. Probing the catalytic mechanism of S-ribosylhomocysteinase (LuxS) with catalytic intermediates and substrate analogues. J. Am. Chem. Soc. 2009, 131, 1243–1250. [Google Scholar] [CrossRef]
  52. Bassler, B.L.; Greenberg, E.P.; Stevens, A.M. Cross-species induction of luminescence in the quorum-sensing bacterium Vibrio harveyi. J. Bacteriol. 1997, 179, 4043–4045. [Google Scholar] [CrossRef]
  53. Lima, E.M.F.; Winans, S.C.; Pinto, U.M. Quorum sensing interference by phenolic compounds—A matter of bacterial misunderstanding. Heliyon 2023, 9, e17657. [Google Scholar] [CrossRef] [PubMed]
  54. Kalia, V.C.; Patel, S.K.S.; Kang, Y.C.; Lee, J.-K. Quorum sensing inhibitors as antipathogens: Biotechnological applications. Biotechnol. Adv. 2019, 37, 68–90. [Google Scholar] [CrossRef] [PubMed]
  55. Truchado, P.; Larrosa, M.; Castro-Ibáñez, I.; Allende, A. Plant food extracts and phytochemicals: Their role as Quorum Sensing Inhibitors. Trends Food Sci. Technol. 2015, 43, 189–204. [Google Scholar] [CrossRef]
  56. Lu, S.Y.; Zhao, Z.; Avillan, J.J.; Liu, J.; Call, D.R. Autoinducer-2 quorum sensing contributes to regulation of microcin PDI in Escherichia coli. Front. Microbiol. 2017, 8, 2570. [Google Scholar] [CrossRef] [PubMed]
  57. De Kievit, T.R.; Gillis, R.; Marx, S.; Brown, C.; Iglewski, B.H. Quorum-sensing genes in Pseudomonas aeruginosa biofilms: Their role and expression patterns. Appl. Environ. Microbiol. 2001, 67, 1865–1873. [Google Scholar] [CrossRef]
  58. Dickschat, J.S. Quorum sensing and bacterial biofilms. Nat. Prod. Rep. 2010, 27, 343–369. [Google Scholar] [CrossRef]
  59. Li, Y.-H.; Tian, X. Quorum sensing and bacterial social interactions in biofilms. Sensors 2012, 12, 2519–2538. [Google Scholar] [CrossRef]
  60. Xavier, K.B.; Bassler, B.L. LuxS quorum sensing: More than just a numbers game. Curr. Opin. Microbiol. 2003, 6, 191–197. [Google Scholar] [CrossRef]
  61. Laganenka, L.; Colin, R.; Sourjik, V. Chemotaxis towards autoinducer 2 mediates autoaggregation in Escherichia coli. Nat. Commun. 2016, 7, 12984. [Google Scholar] [CrossRef]
  62. González Barrios, A.F.; Zuo, R.; Hashimoto, Y.; Yang, L.; Bentley, W.E.; Wood, T.K. Autoinducer 2 controls biofilm formation in Escherichia coli through a novel motility quorum-sensing regulator (MqsR, B3022). J. Bacteriol. 2006, 188, 305–316. [Google Scholar] [CrossRef]
  63. Bansal, T.; Jesudhasan, P.; Pillai, S.; Wood, T.K.; Jayaraman, A. Temporal regulation of enterohemorrhagic Escherichia coli virulence mediated by autoinducer-2. Appl. Microbiol. Biotechnol. 2008, 78, 811–819. [Google Scholar] [CrossRef]
  64. Jani, S.; Seely, A.L.; Peabody, V.G.; Jayaraman, A.; Manson, M.D. Chemotaxis to self-generated AI-2 promotes biofilm formation in Escherichia coli. Microbiology 2017, 163, 1778–1790. [Google Scholar] [CrossRef] [PubMed]
  65. Whiteley, M.; Diggle, S.P.; Greenberg, E.P. Progress in and promise of bacterial quorum sensing research. Nature 2017, 551, 313–320. [Google Scholar] [CrossRef] [PubMed]
  66. Lee, D.J.; Jo, A.R.; Jang, M.C.; Nam, J.; Choi, H.J.; Choi, G.W.; Sung, H.Y.; Bae, H.; Ku, Y.G.; Chi, Y.T. Analysis of two quorum sensing-deficient isolates of Pseudomonas aeruginosa. Microb. Pathog. 2018, 119, 162–169. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, Y.; Qi, K.; Jing, Y.; Zuo, J.; Hu, J.; Lv, X.; Chen, Z.; Mi, R.; Huang, Y.; Yu, S.; et al. LsrB-based and temperature-dependent identification of bacterial AI-2 receptor. AMB Express 2017, 7, 188. [Google Scholar] [CrossRef]
  68. Pereira, C.S.; Thompson, J.A.; Xavier, K.B. AI-2-mediated signalling in bacteria. FEMS Microbiol. Rev. 2013, 37, 156–181. [Google Scholar] [CrossRef]
  69. Papaneophytou, C.; Giannenas, I.; Dragomir, C. Resistance of bacteria, fungi, and parasites to antibiotics or natural substances of botanical origin. In Feed Additives; Florou-Paneri, P., Christaki, E., Giannenas, I., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 339–354. [Google Scholar]
  70. Koh, C.L.; Sam, C.K.; Yin, W.F.; Tan, L.Y.; Krishnan, T.; Chong, Y.M.; Chan, K.G. Plant-derived natural products as sources of anti-quorum sensing compounds. Sensors 2013, 13, 6217–6228. [Google Scholar] [CrossRef]
  71. Meng, F.; Zhang, F.; Chen, Q.; Yang, M.; Yang, Y.; Li, X.; Gu, W.; Yu, J. Virtual screening and in vitro experimental verification of LuxS inhibitors from natural products for Lactobacillus reuteri. Biomed. Pharmacother. 2022, 147, 112521. [Google Scholar] [CrossRef]
  72. Ouyang, J.; Sun, F.; Feng, W.; Sun, Y.; Qiu, X.; Xiong, L.; Liu, Y.; Chen, Y. Quercetin is an effective inhibitor of quorum sensing, biofilm formation and virulence factors in Pseudomonas aeruginosa. J. Appl. Microbiol. 2016, 120, 966–974. [Google Scholar] [CrossRef]
  73. Bouyahya, A.; Chamkhi, I.; Balahbib, A.; Rebezov, M.; Shariati, M.A.; Wilairatana, P.; Mubarak, M.S.; Benali, T.; El Omari, N. Mechanisms, anti-quorum-sensing actions, and clinical trials of medicinal plant bioactive compounds against bacteria: A comprehensive review. Molecules 2022, 27, 1484. [Google Scholar] [CrossRef]
  74. Vikram, A.; Jayaprakasha, G.K.; Jesudhasan, P.R.; Pillai, S.D.; Patil, B.S. Suppression of bacterial cell-cell signalling, biofilm formation and type III secretion system by citrus flavonoids. J. Appl. Microbiol. 2010, 109, 515–527. [Google Scholar] [CrossRef] [PubMed]
  75. Al-Yousef, H.M.; Ahmed, A.F.; Al-Shabib, N.A.; Laeeq, S.; Khan, R.A.; Rehman, M.T.; Alsalme, A.; Al-Ajmi, M.F.; Khan, M.S.; Husain, F.M. Onion peel ethylacetate fraction and its derived constituent quercetin 4′-O-β-D glucopyranoside attenuates quorum sensing regulated virulence and biofilm formation. Front. Microbiol. 2017, 8, 1675. [Google Scholar] [CrossRef]
  76. Erdönmez, D.; Rad, A.Y.; Aksöz, N. Anti-quorum sensing potential of antioxidant quercetin and resveratrol. Braz. Arch. Biol. Technol. 2018, 61, e18160756. [Google Scholar] [CrossRef]
  77. Xu, W.; Zhang, X.; Wang, L.; Zeng, W.; Sun, Y.; Zhou, C.; Zhou, T.; Shen, M. Effect of chlorogenic acid on the quorum-sensing system of clinically isolated multidrug-resistant Pseudomonas aeruginosa. J. Appl. Microbiol. 2022, 132, 1008–1017. [Google Scholar] [CrossRef] [PubMed]
  78. Rama Devi, K.; Srinivasan, R.; Kannappan, A.; Santhakumari, S.; Bhuvaneswari, M.; Rajasekar, P.; Prabhu, N.M.; Veera Ravi, A. In vitro and in vivo efficacy of rosmarinic acid on quorum sensing mediated biofilm formation and virulence factor production in Aeromonas hydrophila. Biofouling 2016, 32, 1171–1183. [Google Scholar] [CrossRef] [PubMed]
  79. Walker, T.S.; Bais, H.P.; Déziel, E.; Schweizer, H.P.; Rahme, L.G.; Fall, R.; Vivanco, J.M. Pseudomonas aeruginosa-plant root interactions. Pathogenicity, biofilm formation, and root exudation. Plant Physiol. 2004, 134, 320–331. [Google Scholar] [CrossRef]
  80. Wang, H.; Chu, W.; Ye, C.; Gaeta, B.; Tao, H.; Wang, M.; Qiu, Z. Chlorogenic acid attenuates virulence factors and pathogenicity of Pseudomonas aeruginosa by regulating quorum sensing. Appl. Microbiol. Biotechnol. 2019, 103, 903–915. [Google Scholar] [CrossRef]
  81. Bouyahya, A.; El Omari, N.; El Menyiy, N.; Guaouguaou, F.-E.; Balahbib, A.; Chamkhi, I. Anti-quorum sensing agents from natural sources. In Antimicrobial Resistance: Underlying Mechanisms and Therapeutic Approaches; Kumar, V., Shriram, V., Paul, A., Thakur, M., Eds.; Springer Nature: Singapore, 2022; pp. 533–557. [Google Scholar]
  82. Nakagawa, S.; Hillebrand, G.G.; Nunez, G. Rosmarinus officinalis L. (rosemary) extracts containing carnosic acid and carnosol are potent quorum sensing inhibitors of Staphylococcus aureus virulence. Antibiotics 2020, 9, 149. [Google Scholar] [CrossRef]
  83. Rasch, M.; Andersen, J.B.; Nielsen, K.F.; Flodgaard, L.R.; Christensen, H.; Givskov, M.; Gram, L. Involvement of bacterial quorum-sensing signals in spoilage of bean sprouts. Appl. Environ. Microbiol. 2005, 71, 3321–3330. [Google Scholar] [CrossRef]
  84. Vattem, D.A.; Mihalik, K.; Crixell, S.H.; McLean, R.J.C. Dietary phytochemicals as quorum sensing inhibitors. Fitoterapia 2007, 78, 302–310. [Google Scholar] [CrossRef]
  85. Yang, M.; Meng, F.; Gu, W.; Li, F.; Tao, Y.; Zhang, Z.; Zhang, F.; Yang, X.; Li, J.; Yu, J. Effects of natural products on bacterial communication and network-quorum sensing. BioMed Res. Int. 2020, 2020, 8638103. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) The predicted 3D structure of E. coli MG1655 LuxS. The enzyme is depicted as a dimer, consisting of two identical monomers shown in red (Chain A) and green (Chain B) ribbon diagrams. Each chain coordinates an Fe2+ ion, represented as grey spheres. (B) The active site of the enzyme, formed by contributions from amino acids of both chains, is highlighted using stick models. The red and green colors indicate Chains A and B, respectively. (C) Multiple sequence alignment of LuxS from E. coli MG1655 (UniProt: P45578) with LuxS proteins from related species, including Haemophilus influenzae (UniProt: P44007), Salmonella typhi (UniProt: Q8Z4D7), Vibrio cholerae (UniProt: Q9KUG4), Campylobacter concisus (UniProt: A7ZGE2), and Neisseria meningitidis (UniProt: A9M0P7). The alignment was performed using Clustal Omega (https://www.ebi.ac.uk/Tools/msa/clustalo/, accessed on 1 August 2024). Arrows indicate key amino acids involved in catalysis (substrate binding or enzyme activity). Identical amino acids are marked with an asterisk (*), conserved amino acids with a colon (:), and semi-conserved amino acids with a period (.).
Figure 1. (A) The predicted 3D structure of E. coli MG1655 LuxS. The enzyme is depicted as a dimer, consisting of two identical monomers shown in red (Chain A) and green (Chain B) ribbon diagrams. Each chain coordinates an Fe2+ ion, represented as grey spheres. (B) The active site of the enzyme, formed by contributions from amino acids of both chains, is highlighted using stick models. The red and green colors indicate Chains A and B, respectively. (C) Multiple sequence alignment of LuxS from E. coli MG1655 (UniProt: P45578) with LuxS proteins from related species, including Haemophilus influenzae (UniProt: P44007), Salmonella typhi (UniProt: Q8Z4D7), Vibrio cholerae (UniProt: Q9KUG4), Campylobacter concisus (UniProt: A7ZGE2), and Neisseria meningitidis (UniProt: A9M0P7). The alignment was performed using Clustal Omega (https://www.ebi.ac.uk/Tools/msa/clustalo/, accessed on 1 August 2024). Arrows indicate key amino acids involved in catalysis (substrate binding or enzyme activity). Identical amino acids are marked with an asterisk (*), conserved amino acids with a colon (:), and semi-conserved amino acids with a period (.).
Macromol 04 00045 g001
Figure 2. Chemical structures of the main components detected in the ethanolic extracts of oregano, rosemary, and common sage from Cyprus flora.
Figure 2. Chemical structures of the main components detected in the ethanolic extracts of oregano, rosemary, and common sage from Cyprus flora.
Macromol 04 00045 g002
Figure 3. Docking studies of the five phytochemicals with the LuxS protein. (A) Overall representation of the binding sites for the five compounds. The LuxS protein is shown in ribbon representation, with Chain A in red and Chain B in green. The phytochemicals are depicted in stick representation with different colors: apigenin (yellow), carnosol (brown), chlorogenic acid (purple), quercetin (cyan), and rosmarinic acid (blue). (B,C) The locations of the five phytochemicals relative to Chain A and Chain B, respectively. All compounds, except quercetin, are positioned within the same active site located in the cavity between the two chains.
Figure 3. Docking studies of the five phytochemicals with the LuxS protein. (A) Overall representation of the binding sites for the five compounds. The LuxS protein is shown in ribbon representation, with Chain A in red and Chain B in green. The phytochemicals are depicted in stick representation with different colors: apigenin (yellow), carnosol (brown), chlorogenic acid (purple), quercetin (cyan), and rosmarinic acid (blue). (B,C) The locations of the five phytochemicals relative to Chain A and Chain B, respectively. All compounds, except quercetin, are positioned within the same active site located in the cavity between the two chains.
Macromol 04 00045 g003
Figure 4. Molecular interactions of LuxS with five phytochemicals, namely apigenin (A), carnosol (B), chlorogenic acid (C), quercetin (D) and rosmarinic acid (E), displayed as 2D images. The binding score (in kcal/mol) for each compound is indicated below their respective interaction diagram. The amino acid residues of LuxS of Chains A and B involved in the interactions are labeled and highlighted in colored circles. These circles correspond to different interactions indicated at the top right panel of this figure. Docking studies were carried out using PyRx v1.1 (AutoDock Vina). The 2D images were obtained using BIOVIA Discovery Studio.
Figure 4. Molecular interactions of LuxS with five phytochemicals, namely apigenin (A), carnosol (B), chlorogenic acid (C), quercetin (D) and rosmarinic acid (E), displayed as 2D images. The binding score (in kcal/mol) for each compound is indicated below their respective interaction diagram. The amino acid residues of LuxS of Chains A and B involved in the interactions are labeled and highlighted in colored circles. These circles correspond to different interactions indicated at the top right panel of this figure. Docking studies were carried out using PyRx v1.1 (AutoDock Vina). The 2D images were obtained using BIOVIA Discovery Studio.
Macromol 04 00045 g004
Figure 5. Inhibition of AI-2 signaling by apigenin (A), carnosol (B), chlorogenic acid (C), quercetin (D), and rosmarinic acid (E). The percentage of AI-2 inhibition is plotted against the concentration of each compound (ranging from 0 to 200 μM). Each graph also includes the chemical structure of the corresponding phytochemical. The IC50 values indicate the concentration of each compound required to inhibit 50% of AI-2 signaling, highlighting differences in their inhibitory potencies.
Figure 5. Inhibition of AI-2 signaling by apigenin (A), carnosol (B), chlorogenic acid (C), quercetin (D), and rosmarinic acid (E). The percentage of AI-2 inhibition is plotted against the concentration of each compound (ranging from 0 to 200 μM). Each graph also includes the chemical structure of the corresponding phytochemical. The IC50 values indicate the concentration of each compound required to inhibit 50% of AI-2 signaling, highlighting differences in their inhibitory potencies.
Macromol 04 00045 g005
Figure 6. Effects of five phytochemicals—apigenin, carnosol, chlorogenic acid (chlor. acid), quercetin (querc.), and rosmarinic acid (rosm. acid)—on biofilm formation by E. coli MG1655, as quantified by crystal violet staining and measuring absorbance at 570 nm. Data are presented as the percentage inhibition of biofilm formation compared with the control (no compound). Results are shown as mean values ± standard deviation from three independent experiments. Error bars represent standard deviations. Statistical analysis was performed using ANOVA followed by Tukey’s multiple-comparison test. Only statistically significant differences are shown and indicated by asterisks: * p < 0.05, ** p < 0.01, *** p < 0.001.
Figure 6. Effects of five phytochemicals—apigenin, carnosol, chlorogenic acid (chlor. acid), quercetin (querc.), and rosmarinic acid (rosm. acid)—on biofilm formation by E. coli MG1655, as quantified by crystal violet staining and measuring absorbance at 570 nm. Data are presented as the percentage inhibition of biofilm formation compared with the control (no compound). Results are shown as mean values ± standard deviation from three independent experiments. Error bars represent standard deviations. Statistical analysis was performed using ANOVA followed by Tukey’s multiple-comparison test. Only statistically significant differences are shown and indicated by asterisks: * p < 0.05, ** p < 0.01, *** p < 0.001.
Macromol 04 00045 g006
Figure 7. Effects of the phytochemicals apigenin, carnosol, chlorogenic acid (chlor. acid), quercetin, and rosmarinic acid (rosm. acid) on E. coli MG1655 growth. (A) E. coli MG1655 was incubated in the absence and presence of 200 μM of each phytochemical in a 96-well plate (200 μL per well) for 20 h. (B) The effects of the phytochemicals at a final concentration of 200 μM on bacterial growth were also determined using viable plate counts after 24 h of incubation, as described in the text. Results are presented as mean values ± SD, with n = 3.
Figure 7. Effects of the phytochemicals apigenin, carnosol, chlorogenic acid (chlor. acid), quercetin, and rosmarinic acid (rosm. acid) on E. coli MG1655 growth. (A) E. coli MG1655 was incubated in the absence and presence of 200 μM of each phytochemical in a 96-well plate (200 μL per well) for 20 h. (B) The effects of the phytochemicals at a final concentration of 200 μM on bacterial growth were also determined using viable plate counts after 24 h of incubation, as described in the text. Results are presented as mean values ± SD, with n = 3.
Macromol 04 00045 g007
Table 1. Properties and binding scores of the five phytochemicals against LuxS.
Table 1. Properties and binding scores of the five phytochemicals against LuxS.
Compound PubChem
CID
Molecular Mass (g/mole)cLogP aBinding Score b (kcal/mol)Main Binding
Amino Acid Residues c
Chain AChain B
Apigenin5280443270.242.90−7.2His54, Glu57, His58Ser6, Phe7, His11, Cys83
Carnosol442009330.403.16−7.0His54, Glu57, His58, Cys128Ser6, His11, Arg39, Cys83,
Chlorogenic acid1794427354.31−1.88−7.3His54, Glu57, His58, Cys128Ser6, Phe7, His11, Cys83
Quercetin5280343302.231.50−6.7--
Rosmarinic acid5281792360.301.10−8.3Glu57, His58, Cys128Ser6, Phe7, His11, Arg39
a Calculated with ChemDraw v18.2; b Calculated with PyRx v1.1 (AutoDock Vina v.1; https://vina.scripps.edu/, accessed on 2 August 2024) as described in the text; c While several amino acids participate in substrate binding, only conserved residues are illustrated. All contributing amino acids involved in the binding of each compound are shown in Figure 4.
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

Sarigiannis, Y.; Papaneophytou, C. Targeting Bacterial Communication: Evaluating Phytochemicals as LuxS Inhibitors to Disrupt Quorum Sensing. Macromol 2024, 4, 753-771. https://doi.org/10.3390/macromol4040045

AMA Style

Sarigiannis Y, Papaneophytou C. Targeting Bacterial Communication: Evaluating Phytochemicals as LuxS Inhibitors to Disrupt Quorum Sensing. Macromol. 2024; 4(4):753-771. https://doi.org/10.3390/macromol4040045

Chicago/Turabian Style

Sarigiannis, Yiannis, and Christos Papaneophytou. 2024. "Targeting Bacterial Communication: Evaluating Phytochemicals as LuxS Inhibitors to Disrupt Quorum Sensing" Macromol 4, no. 4: 753-771. https://doi.org/10.3390/macromol4040045

APA Style

Sarigiannis, Y., & Papaneophytou, C. (2024). Targeting Bacterial Communication: Evaluating Phytochemicals as LuxS Inhibitors to Disrupt Quorum Sensing. Macromol, 4(4), 753-771. https://doi.org/10.3390/macromol4040045

Article Metrics

Back to TopTop