Next Article in Journal
Flexible Coatings Facilitate pH-Targeted Drug Release via Self-Unfolding Foils: Applications for Oral Drug Delivery
Next Article in Special Issue
LC-AMP-F1 Derived from the Venom of the Wolf Spider Lycosa coelestis, Exhibits Antimicrobial and Antibiofilm Activities
Previous Article in Journal
Curcumin for Treating Breast Cancer: A Review of Molecular Mechanisms, Combinations with Anticancer Drugs, and Nanosystems
Previous Article in Special Issue
Lactobacillus johnsonii LJO02 (DSM 33828) Cell-Free Supernatant and Vitamin D Improve Wound Healing and Reduce Interleukin-6 Production in Staphylococcus aureus-Infected Human Keratinocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Shifting from Ammonium to Phosphonium Salts: A Promising Strategy to Develop Next-Generation Weapons against Biofilms

Department of Pharmacy, University of Genoa, Viale Cembrano, 4, 16148 Genova, Italy
Submission received: 24 November 2023 / Revised: 3 January 2024 / Accepted: 4 January 2024 / Published: 5 January 2024

Abstract

:
Since they are difficult and sometimes impossible to treat, infections sustained by multidrug-resistant (MDR) pathogens, emerging especially in nosocomial environments, are an increasing global public health concern, translating into high mortality and healthcare costs. In addition to having acquired intrinsic abilities to resist available antibiotic treatments, MDR bacteria can transmit genetic material encoding for resistance to non-mutated bacteria, thus strongly decreasing the number of available effective antibiotics. Moreover, several pathogens develop resistance by forming biofilms (BFs), a safe and antibiotic-resistant home for microorganisms. BFs are made of well-organized bacterial communities, encased and protected in a self-produced extracellular polymeric matrix, which impedes antibiotics’ ability to reach bacteria, thus causing them to lose efficacy. By adhering to living or abiotic surfaces in healthcare settings, especially in intensive care units where immunocompromised older patients with several comorbidities are hospitalized BFs cause the onset of difficult-to-eradicate infections. In this context, recent studies have demonstrated that quaternary ammonium compounds (QACs), acting as membrane disruptors and initially with a low tendency to develop resistance, have demonstrated anti-BF potentialities. However, a paucity of innovation in this space has driven the emergence of QAC resistance. More recently, quaternary phosphonium salts (QPSs), including tri-phenyl alkyl phosphonium derivatives, achievable by easy one-step reactions and well known as intermediates of the Wittig reaction, have shown promising anti-BF effects in vitro. Here, after an overview of pathogen resistance, BFs, and QACs, we have reviewed the QPSs developed and assayed to this end, so far. Finally, the synthetic strategies used to prepare QPSs have also been provided and discussed to spur the synthesis of novel compounds of this class. We think that the extension of the knowledge about these materials by this review could be a successful approach to finding effective weapons for treating chronic infections and device-associated diseases sustained by BF-producing MDR bacteria.

1. Antimicrobial Resistance

Antibiotics have been an essential tool of modern medicine to fight illnesses, but antibiotic treatment is becoming increasingly unreliable as antimicrobial resistance (AMR) has evolved in many bacterial pathogens [1]. Antimicrobial resistance happens when germs like bacteria, fungi, and viruses develop the ability to defeat the drugs designed to kill them [2]. Despite the use of antibiotics initially functioning, germs that have developed resistance continue to grow, making infections increasingly difficult, and sometimes impossible, to treat [3]. Antimicrobial resistance is a naturally occurring process promoted mainly by the misuse and overuse of antimicrobials, thus causing the need for costly new treatments to be developed [4]. Antimicrobial-resistant organisms continue to take thousands of lives per year, including a wide variety of people, such as those within the medical community, public health officials, food producers, pharmaceutical developers, and the general public. Antimicrobial resistance is an urgent global public health threat, with the forecast of 10 million deaths per year globally by 2050 [4]. Gram-negative bacteria, such as Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, Burkholderia cepacian, and Escherichia coli, are responsible for severe infections including pneumonia, bloodstream infections, wound or surgical site infections, and meningitis in healthcare settings and especially in intensive care units where immunocompromised older patients with several comorbidities are hospitalized [5]. Since antibiotics from the carbapenem family are now the last line of defense against antibiotic-resistant bacterial infections, carbapenem-resistant Enterobacteriaceae, extensively drug-resistant (XDR) P. aeruginosa and XDR A. baumannii are the most clinically relevant resistance phenotypes [5].
Antibiotic degradation, antibiotic target modification, modulation of membrane permeability, structural modifications of bacterial lipopolysaccharide, hyper-expression of efflux pumps, as well as the formation of biofilms (BFs) are some of the recognized mechanisms used by bacteria to develop resistance [6]. Table S1 in the Supplementary Materials reports some of the most clinically relevant MDR pathogens with their associated infections [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38].
To ameliorate this worrying scenario, where antibacterial resistance is incessantly increasing, the available effective antibiotics are decreasing dramatically, and infections sustained by BF-producing pathogens are practically untreatable, researchers have focused mainly on developing drugs acting on contact as membrane disruptors that therefore have a low tendency to develop resistance. In this context, quaternary ammonium salts (QASs) have demonstrated promising biocidal effects and anti-BF potentialities [39,40,41,42,43,44]. Unfortunately, their continuous use over more than 80 years has driven the emergence of QAS resistance [45]. In this regard, quaternary phosphonium salts (QPSs), including tri-phenyl alkyl phosphonium derivatives, achievable by easy, low-cost, one-step reactions and already known as intermediates of the Wittig reaction, have shown promising biocidal and anti-BF effects in vitro. In this paper, to draw more attention to QPSs, which are still too little studied by researchers, after having provided essential information on pathogen resistance and BFs and a discussion on QASs, we have reviewed the QPSs developed and assayed to this end in recent years. Finally, to spur the synthesis of novel compounds of this class, potentially active against BFs, the synthetic strategies used to prepare the antibacterial QPSs developed so far have been also provided and discussed.

2. Biofilms: The Antibiotic-Resistant Home Self-Produced by Microorganisms

A particularly worrying form of pathogen resistance is represented by biofilms (BFs), which are a safe and antibiotic-resistant home to microorganisms [46]. BFs are well-organized heterogeneous communities of self-immobilized microorganisms, attached to each other and to a surface, which are encased within self-produced extracellular polymeric substances (EPSs). EPSs protect them from environmental stressors, including extreme conditions of pH, oxygen, osmotic shock, heat, freezing, UV radiation, predators, disinfectants, and antibiotics [47]. In BFs, cell–cell interactions are facilitated, including those that allow the transfer of resistance genes, and microorganisms cooperate in a synchronized way to capture resources, such as nutrients [7,48]. BFs are a lifestyle for 40–80% of prokaryotic or unicellular eukaryotic cells [49], representing an incubator for genotypic and phenotypic variants [48]. In fact, BFs should be regarded as a multicellular superorganism where sessile cells, persistent cells, and dormant cells, with physiological characteristics differentiating them from planktonic cells, cohabit and communicate by the quorum sensing (QS) system [7,48]. The extracellular polymeric matrix in which BF-producing cells are encased consists of extracellular polymeric substances (EPSs), including a combination of enzymatic proteins, polysaccharides (cellulose, poly-glucosamine (PGA), and exopolysaccharides), extracellular DNA (eDNA), as well as cationic and anionic glycoproteins and glycolipids, which allow real communication between the bacteria and stabilize the three-dimensional structure of the BF itself [50,51]. Table S2 in the Supplementary Materials schematizes the general composition of BF biomass.
EPSs not only protect bacteria from external dangers but also provide them with the necessary nourishments and water, which is retained by means of H bonds with hydrophilic polysaccharides. Depending on the required nutrients, the microorganisms in a BF can change the composition of the EPSs by secreting proper signaling molecules [50,51]. Although beneficial microorganisms could form BFs associated with human, animal, and plant hosts as an essential part of the holobiont [52,53], BFs produced by pathogens are responsible for about 80% of bacterial infections [54], which are often extremely difficult to treat due to the specific protection mechanisms provided by the BF to bacterial cells, immobilized within the EPSs and not reachable by antibiotics [55]. Indeed, BF-producing cells are 10–1000-fold less susceptible to various antimicrobial agents than their planktonic forms. BF-related infections can be divided into device-associated infections and non-device-associated ones. Table 1 collects the most relevant BF-producing pathogens and their related infections.

2.1. Biofilm Formation and Dispersal

Bacteria can exhibit a free-living (mobile) planktonic state of living and a sessile (non-mobile) biofilm-living state when adhered to a substrate [72]. The in vitro BF life cycle involves the different stages reported in Table S3 in the Supplementary Materials [73] and visualized in Figure 1.
At stage 5 concerning the detachment and dispersion of cells from the BF, the initiation of new BF formation has been observed. Dispersed cells are morphologically more like planktonic cells than those of mature BFs and are crucial for developing into a new BF [61]. Reproducible periodic BF dispersal in bacterial species such as Actinobacillus spp., Actinomycetes comitans [75], P. aeruginosa [76], and Serratia marcescens [77] has been reported. The main factors, cues, and signals that induce BF detachment include temperature, pH, enzymatic degradation, starvation or excess of nutrients, presence of D-amino acids, nitric oxide, biosurfactants, and QS signaling molecules. In vitro and in vivo experiments have shown that the cells liberated by BFs are extremely cytotoxic to macrophages, more sensitive to iron diminution, and significantly more virulent to nematode hosts than planktonic bacteria. Additionally, it was reported that dispersed bacteria derived from BFs treated with glycoside hydrolase rapidly induced fatal septicemia in a mouse chronic wound infection model.

2.1.1. Quorum Sensing (QS) in Bacterial Biofilms

The formation of BF is governed by physiological processes and by the coordinated activities of the population of germs, which communicate with each other mainly through the quorum sensing (QS) machine [78]. Particularly, QS is a specific intercellular communication method based on the release of signal molecules that diffuse in BFs and are detected by other microorganisms, which in turn respond. QS machinery plays a vital role in BF formation, exopolysaccharide synthesis, motility, and chemotaxis, which are imperative for bacteria during the degradation or detoxification of any pollutant [78]. QS, first reported in the marine bioluminescent bacterium Vibrio fischeri [79], is currently the paramount target of strategies to control BF formation by disrupting cell-to-cell communication, conjugation, nutrient acquisition, and even motility and production of certain metabolites [78]. QS signaling molecules are different in Gram-positive and Gram-negative bacteria. There are at least three types of QS systems, including the acyl-homoserine lactone system (AHL) in Gram-negative bacteria, the autoinducing oligopeptide (API) QS system in Gram-positive bacteria, and the LuxS-encoded autoinducer-2 QS system (AI-2) in both species [80]. AHL and API QS systems, which are based on signaling molecules such as acyl-homoserine lactones and small peptides, respectively, are planned specifically for “intraspecies” signaling. An increase in AHL molecules corresponds to high bacterial growth [81], while high concentrations of APIs affect target gene expression, which are responsible for the production of numerous toxins and degradable exoenzymes [78]. The production of AI-2 signals, also called “universal language”, is catalyzed by LuxS synthase and allows communication between microorganisms of different species [82]. Moreover, LuxS is involved in the activation of the methylation cycle, which is involved in the control of the expression of hundreds of genes associated with the microbial processes of surface adhesion, detachment, and toxin production [82].

2.1.2. Cyclic Dimeric Guanosine Monophosphate (c-di-GMP) in Bacterial Biofilms

Bis-cyclic dimeric guanosine monophosphate (c-di-GMP) is a secondary messenger that mediates crucial cellular activities in bacteria, including growth, motility, virulence, BF formation, and cell cycle progression [83]. In particular, c-di-GMP is involved in the transformation of planktonic cells to sessile ones during BF formation, as well as in the passage from the sessile lifestyle to the planktonic one through the BF dispersal process [84]. The BF formation and dispersal processes are governed by c-di-GMP through several genetic interactions. In Gram-negative bacteria, low and high c-di-GMP levels correlate with the motile and sessile phenotypes, respectively [83], so high concentrations lead to BF formation, while low concentrations stimulate motility and thus dispersal of bacterial cells from BF [84]. Such a correlation has been demonstrated especially in E. coli, P. aeruginosa, and Salmonella typhimurium [85].

3. Current and New Therapeutic Approaches against BF

Infections caused by BFs are a worrying problem causing chronic diseases that are difficult to treat and require high-dose antibiotics according to the severity of infection. As previously reported, cells in BFs are up to 1000-fold more tolerant to antibiotics and disinfectants than planktonic ones [86]. Additionally, when infections are device-associated and antibiotics fail due to the increasing resistance of bacteria in BFs, surgical replacement of the infected devices is necessary [87]. In this regard, it has been reported that the best possible treatment for BF-based infections could be the possibility of an early diagnosis using biosensors, advanced imaging, or theranostic nanoparticles (NPs), as well as inhibiting the initial attachment stage, thus preventing the infection from starting [86]. Innovative and effective antibiotic strategies have been designed to counteract BF formation and promote its dispersal, such as treatments for dispersing BFs by dissolution of EPSs, associations of conventional antibiotics with QS inhibitors, and a mixture of these novel techniques. Further advanced treatments include antimicrobial photodynamic therapies, bioacoustics, as well as electric and magnetic fields [88]. However, this research area is still too little explored and far from undergoing clinical research and entering the commercial market [88]. Currently, there are a few molecules under clinical development that are active against MDR pathogens and/or BF producers. Among the molecules in Phase III clinical trials in 2020, only caspofungin, which is an antifungal drug, has been demonstrated to inhibit the synthesis of the polysaccharide components of the S. aureus BF [89]. In the field of NPs, polymeric lipid NPs deriving from the conjugation of rhamnolipids (biosurfactants secreted by P. aeruginosa) and polymer NPs made of clarithromycin encapsulated in a polymeric core of chitosan have been studied, while rhamnolipid-coated silver and iron oxide NPs have been developed, which were effective in eradicating S. aureus and P. aeruginosa BFs [89]. Naturally occurring cationic antimicrobial peptides (CAMPs), a class of non-β-lactam antimicrobial agents, acting on contact as membrane disruptors, have been demonstrated to be effective on a wide variety of Gram-positive and Gram-negative bacteria, fungi, protozoa, and yeast and to be active alone or in combination with conventional antibiotics against Staphylococci and P. aeruginosa BFs [90,91,92,93]. So, inspired by CAMPs, several cationic compounds have been developed that demonstrate similar mechanisms of action and effects. In this context, some antimicrobial cationic dendrimers have proven anti-BF effects and the capability to penetrate EPSs and prevent BF formation, thus representing novel promising substances to counteract this alarming form of bacterial resistance [7,88]. Quaternary ammonium salts (QASs) include the most recognizable commercial disinfectants and antiseptics, such as benzalkonium chloride [39], miramistin [40], di-decyl-dimethyl-ammonium chloride [41], cetyl-pyridinium chloride [42], and octenidine dichloride [43]. Moreover, studies reported that surfactants classified as gemini quaternary ammonium salts (g-QASs) have been inserted in biocidal preparations which were active against the biofilm created by P. aeruginosa by disrupting its functioning, and lysing BF cells upon electrostatic interactions [44]. In the case of BFs formed by Gram-positive bacteria, g-QASs such as alanine gemini surfactants manifested anti-BF activity against S. epidermidis [44]. With the aim of preventing the phenomenon of adhesion of microorganisms, surfaces modified with QASs and their gemini derivatives have been developed. Moreover, the ability of QASs to eradicate C. albicans and R. mucilaginosa BFs was observed [44].

3.1. Nitrogen-Based Quaternary Salts

Permanently cationic nitrogen-based surfactants, such as quaternary ammonium salts (QASs) and derivatives are made of one or more polar cationic heads and differently structured hydrophobic carbon tails, often including long carbon chains. QASs possess algistatic, bacteriostatic, tuberculostatic, sporostatic, fungistatic, and antiviral activities [94,95,96,97]. Although their real mechanism of action has not been fully understood, as schematically shown in Figure 2, it has been reported that at a concentration close to the MIC, after the initial adsorption of QASs on the cell wall of pathogens by electrostatic interaction, they penetrate it, due to the presence of hydrophobic alkyl chains [98].
Here, QASs react with lipids and proteins of the cell membrane, thus causing disorganization in its structure, pore formation, loss of osmoregulatory and physiological functions, and leakage of low-molecular-weight components out of the cell [98]. Subsequently, proteins and nucleic acids degrade inside the cell, and the release of autolytic enzymes occurs, which leads to the lysis of the cell wall components and a complete loss of the structural organization of the cell [98]. At concentrations higher than the MIC, aggregates are formed that solubilize hydrophobic membrane elements [99]. It has been shown that the highest biocidal activity against Gram-positive bacteria and yeast was provided by QASs with C10–C14 carbon chains, while QASs with C14–C16 carbon chains were biocidal against Gram-negative bacteria [100]. However, to exert their biocidal effects, QASs need to pass through the outer envelope of the bacterial cells. In this regard, since that of Gram-negative species is composed of more layers than that of Gram-positive strains, it is not surprising that Gram-negative bacteria show more resistance to QASs than Gram-positive species [101]. Concerning BFs, the inhibition of microbial adhesion to surfaces seems to be a good target for their prevention, and numerous studies have focused on surface modification in order to impede the interaction of microbial cells with it. In this context, the immobilization of QASs on surfaces has been reported as a successful way of reducing bacterial cell attachment. In the past years, the groups of Tan and Peng have shown that chitosan derivatives of monomeric QASs deposited on bone cement and titanium inhibited the adhesion of Staphylococci [102,103], while their incorporation into silica NPs specifically reduced the adhesion of S. aureus [104]. Although the anti-adhesion effect of QASs also depends on the material properties of the surfaces (hydrophobicity, charge) and on the characteristics of the cell envelope, it has been demonstrated that the presence of the second QAS molecule in the gemini structures (Figure 3) enhances their surface activity, resulting in a stronger interaction with the material [105].
Concerning this, polystyrene coatings decorated with glycine and alanine gemini QASs reduced the adhesion of S. epidermidis and C. albicans, with the best effects achieved for C12 and C14 compounds. A greater challenge is represented by eradicating mature BFs due to the presence of EPSs and sessile cells with low metabolic activity. Glycine- and alanine-derived gemini QASs with 12 carbons in the chains exhibited not only high antimicrobial effects against planktonic forms of bacteria and fungi but also impressive biofilm-eradicating properties. Specifically, gemini-like surfactants were able to penetrate the BF structure by extracting adherent cells or lowering the viability of cells within the BF. In 2013, Obłak et al. synthesized a series of cationic gemini chlorides and bromides, which poorly reduced the adhesion of microorganisms to the polystyrene plate. However, one compound eradicated the BFs of C. albicans and Rodotorula mucilaginosa, thus resulting in promise in overcoming catheter-associated infections [106]. Later, Tran et al. developed polyurethane (PU) foam wound dressings coated with poly diallyl-dimethylammonium chloride (pDADMAC–PU), which was able to inhibit the growth and development of BFs of S. aureus, P. aeruginosa, and A. baumannii within the wound dressing. A colony-forming unit assay revealed that the pDADMAC–PU dressing produced more than an eight-log reduction in the BF formation of each pathogen [107]. More recent examples of the nitrogen-based permanently cationic materials studied as anti-BF agents have been included in Table 2, which collects mainly the foremost compounds developed in the last five years.

3.2. Quaternary Phosphonium Salts (QPS)

Although they are thermally and chemically more stable [154,155,156,157,158,159] and often display increased antimicrobial properties and lower toxicity than their nitrogen-based counterparts [155,160,161,162], phosphonium compounds as antibacterial agents are inexplicably lesser reported than ammonium ones. Chart 1 reports the chemical structures of some QPSs developed in recent years that have shown remarkable antibacterial and anti-biofilm effects.
Nevertheless, compounds with phosphonium moieties have been reported and used in various biomedical applications, such as antibacterials [163], anticancer agents [164], and water treatment [165,166]. Concerning their mechanism of action, like CAMPs and QASs, QPSs also act on contact as membrane disruptors, following the phases reported in Figure 2. However, it has recently been demonstrated that some representatives of this class, including alkyl triphenyl phosphonium and alkyl tributyl phosphonium salts, are also endowed with remarkable antioxidant activity, thus being capable of reducing oxidative stress triggered by bacteria, helping the human body in fighting them and blocking the replication of microbial pathogens [158]. In the past, the antioxidant properties of tetrakys (THPS) and its biocidal activity by disrupting the disulfide bonds in bacterial cells were reported by Keasler et al. (2010) [149]. The remarkable antimicrobial effects of phosphonium ionic liquids (PILs) have been reported by Brunel et al. [167]. Benzyl triphenyl phosphonium salts exhibited anti-trypanosomal activity against Trypanosoma brucei, and it was established that the existence of bulky substituents (alkyl or aryl) surrounding the phosphonium ions resulted in a reduction in the MIC of these ILs [168]. Bis-phosphonium salt-derived benzophenone exhibited remarkable toxicity to the human protozoan parasite Leishmania [169]. More recently, Das et al. reported the development of mono- and bis-PILs through a purely ionic approach, and the resulting PILs demonstrated selective bacterial toxicity toward Gram-positive S. aureus and Gram-negative E. coli depending on the number of phosphonium ions present in the salt [154]. Phosphonium salts with single and double alkyl chains were investigated by Endo and coworkers [160]. Herein, the effect of the chain length (C10–C18) on antibacterial activity against a wide range of pathogens including MRSA was examined. The study brought forward a direct correlation between molecular structure and antibacterial activity. In particular, it was observed that higher alkyl chain lengths and the presence of aryl groups improved the antimicrobial activity of PILs. However, the aforementioned studies reported the specific use of phosphonium salts as biocides, while studies conducted on their potential activity against BFs are very limited. In this regard, Kim et al. reported on the antifouling properties of tributyl tetradecyl phosphonium chlorides in reverse osmosis processes, which were capable of reducing the thickness and volume of a BF of P. aeruginosa [170]. Among QPSs, triphenyl alkyl phosphonium compounds, known for years as chemical intermediates of the Wittig reaction, easily synthesizable by a one-step low-cost reaction and easily purifiable by crystallizations, have been re-evaluated as promising template molecules to prepare new antibacterial agents active against bacterial BFs. In this regard, a report by Fernández and co-workers described several bioeffects of numerous alkyl-triphenyl phosphonium salts, including the broad-spectrum biocidal activity of compounds with alkyl chains > C7, the antifouling properties of some of them, and the abilities of some others to act as non-biocidal, non-toxic QS disruptors, thus without tendency to develop resistance in microorganisms [171]. In this regard, Melander and co-workers pointed out that it would be extremely important and desirable to develop anti-biofilm agents operating via non-biocidal mechanisms for several reasons, the most important one being avoiding resistance development [172]. In this context, Joseph et al. found that while ammonium and methyl imidazolium cationic pillar[n]arenes were effective inhibitors of BF formation in several strains of Gram-positive bacteria, they were deprived of any antimicrobial activity, thus not causing damage to red blood cells or toxicity to human cells in culture [152]. In the same year, the same authors prepared both ammonium and phosphonium derivatives, which demonstrated similar anti-biofilm properties [153]. Collectively, to have an idea of the discrepancy that exists between the interest in QASs rather than QPSs as possible anti-biofilm agents, Figure 4 shows how many experimental publications have been developed in the last 15 years reporting on quaternary ammonium salts (602) vs. those reporting on quaternary phosphonium salts (13) that were demonstrated to be effective against BFs and BF-based biofouling according to Scopus and PubMed.
Unfortunately, the paucity of innovation evidenced in Figure 4 concerning “onium” compounds has driven the emergence of QAS resistance, and nowadays it is necessary to identify the next generation of disinfectant molecules with efficacy against highly resistant BF-producing clinical isolates [45]. In this regard, a more intensive study on phosphonium-containing compounds could be a rational route to achieve next-generation disinfectants. As confirmation of this thesis, Michaud et al. recently tested a series of quaternary ammonium and quaternary phosphonium compounds (QPCs) against a panel of 35 resistant A. baumannii clinical isolates. The authors found that in contrast to QASs, which developed resistance, a QPC, namely P6P-10,10, maintained efficacy against resistant strains with an IC90 of 3 µM and an MBEC as low as 32 µM against extensively drug-resistant clinical isolates [45]. Furthermore, this year, Shi et al. reported on the synthesis, characterization, and microbiological evaluation of alkyl-bis-(triphenyl) phosphonium bromides, namely (1,2-DBTPP)Br2, (1,4-DBTPP)Br2, and (1,6-DBTPP)Br2. The biological results revealed that the butyl derivative (1,4-DBTPP)Br2 exhibited low toxicity and low hemolytic activity on eukaryotic cells and the capability to inhibit bacterial growth and BF formation and to promote the recovery process of infected wounds in vivo [173].
In addition to those discussed here, the QPSs developed and studied in the last 15 years have been included in Table 3.

4. Synthetic Strategies Applied to Prepare the Aforementioned QPSs

In this last section, the synthetic procedures followed to synthesize the best-performing QPSs included in Table 3 have been reported, providing reaction Schemes, molecule structures, and some more detail.

4.1. Synthesis of Chitosan-Based QPSs

Chitosan QPSs were prepared by means of a two-step reaction. The first step consisted of the synthesis of quaternary benzyl phosphonium salts with a p-carboxylic acid group, which was used to graft them onto commercial chitosan, thus obtaining the related chitosan-based QPS derivatives (Scheme 1).
Particularly, 4-(chloromethyl) benzoic acid was dissolved in ethyl acetate (EtOAc) and treated with trimethyl phosphine or triphenylphosphine under constant magnetic stirring. The mixture was heated to 85 °C for 10 h, then cooled to room temperature. The precipitate was collected by filtration and subjected to repeated washing with EtOAc to afford the desired products, which were dried in a vacuum at 60 °C for 24 h.

4.2. Synthesis of Tetrakis (Hydroxymethyl) Phosphonium Salt (THPS)

A very effective method of preparing tetrakis (hydroxymethyl) phosphonium salt (THPS) consists of reacting phosphine (PH3) with excess formaldehyde in the presence of hydrogen chloride, according to Scheme 2.

4.3. Synthesis of Bis-Phosphonium Salts of Pyridoxine

The bis-phosphonium salts reported by Kayumov et al. (namely compounds 4, 5, and 6 in their paper) can be prepared following the procedure described and then optimized by Pugachev et al. [178,179]. Briefly, the chlorine derivatives of pyridoxine 1 (compounds 2 and 3) were prepared by first introducing the ketal or acetal protections using acetone or butyl aldehyde, respectively, in acidic conditions. Then, an additional hydroxymethyl group was introduced into position 6 of the pyridine ring of pyridoxine by hydroxy methylation in an alkaline medium. Finally, by treating the hydroxyl derivatives with SOCl2 at room temperature, 2 and 3 were achieved, which were treated with a 2.0-fold molar excess of triphenyl or tributyl phosphine in CH3CN at reflux temperature for 7 h, achieving compounds 4, 6, and 5, respectively (Scheme 3).

4.4. Synthesis of Pillar [5]Arene-Based QPSs

The pillar [5]arene-based QPSs 1 and 2 and monomer (M) reported by Joseph et al. [153] were prepared according to Scheme 4.
Briefly, compound 1a was prepared by refluxing hydroquinone and a strong excess of 1,3-dibromopropane in the presence of potassium carbonate (K2CO3) in acetone for 24 h under an argon atmosphere. The reaction mixture was then cooled to 25 °C and filtered through celite, the solvent was evaporated under vacuum, and the residue was dissolved in dichloromethane, washed with water, 3 N HCl, and brine, dried with sodium sulphate, and concentrated in vacuo. After purification by column chromatography and recrystallization, the obtained white solid was converted into compound 1b by dissolution in 1,2-dichloroethane and treatment with paraformaldehyde followed by BF3·OEt2 at 30 °C for 30 min under an argon atmosphere. Compound 1b was obtained as a white solid after precipitation in methanol and purification by chromatography. Finally, by refluxing 1b in a pressure tube in acetonitrile with trimethyl phosphine or triethyl phosphine dissolved in tetrahydrofuran (THF) for 72–96 h, compound 1 or 2 was obtained as white solids in an 86% or 41% yield. Monomer M was prepared as compound 1 but starting from compound 1a rather than 1b.

4.5. Synthesis of Tributyl Tetradecyl Phosphonium Chloride (TTPC)

Tributyl tetradecyl phosphonium chloride (TTPC) was easily synthesized following the procedure proposed by Swapnil Dharaskar and Mika Sillanpaa [180] (Scheme 5).
Briefly, tri-butyl-phosphine was added to an equimolar amount of 1-chlorotetradecane at 140 °C under a nitrogen (N2) atmosphere and was vigorously stirred for 12–16 h. After reaction completion, the mixture was vacuum stripped to remove volatile organic components and excess 1-chlorotetradecane. TTPC was obtained as a clear, pale-yellow liquid.

4.6. General Synthesis of Alkyl Triphenyl Phosphonium Bromide (ATPB)

The most common method to prepare the alkyl triphenyl phosphonium salts used by Martín-Rodríguez et al. in their study [171] involves the quaternization of triphenyl phosphine with the proper alkyl halides at 140 °C, according to Scheme 6 [181].

4.7. Synthesis of Tetra Alkyl Tetraphenyl Bis-Phosphonium Compound (TATPBP, Namely P6P-10,10)

The synthesis of the best-performing biocidal QPS reported by Michaud et al. [45], was previously reported by the same authors (Scheme 7) [182].
Briefly, to 1,6-bis (diphenyl phosphino) hexane (1) in acetonitrile, 1-bromodecane was added. The obtained solution was heated to reflux and stirred for 24 h. After cooling to room temperature and solvent removal, an oil was obtained, which was triturated with 1/1 ether/hexane and cooled at −25 °C overnight. The resulting precipitate was dissolved in dichloromethane and then concentrated to afford P6P-10,10 as a white crystalline powder, in a 94% yield.

4.8. Synthesis of Alkyl Triphenyl Bis-Phosphonium Bromides

A simple one-step reaction of triphenylphosphine with proper dibromo alkanes was carried out to obtain the alkyl-bis-(triphenyl) phosphonium bromides (1,2-DBTPP)Br2, (1,4-DBTPP)Br2, and (1,6-DBTPP)Br2 (Scheme 8).
Briefly, 1,2-, 1,4-, and 1,6-haloalkanes and Ph3P were dissolved in DMF and were reacted at 130 °C for 4 h. The mixtures were then cooled to room temperature, and the solvent was removed by rotation volatilization to extract the raw products, which were washed with n-hexane and dried in an oven.

4.9. Synthesis of N-Phosphonium Chitosans (NPCSs) with Different Degrees of Substitution

N-phosphonium chitosans (NPCSs) with different degrees of substitution (3%, 13%, and 21%) were synthesized according to previous studies by the same authors [183,184] (Scheme 9).
In brief, chitosan and hydroxybenzotriazole (HOBt) were stirred in an H2O/DMSO (v/v = 2/1) mixture overnight at 15 °C until the homogeneous solution was obtained. Then, commercial CTPB dissolved in H2O/DMSO (v/v = 2/1) was added to the solution, followed by the dropwise addition of a solution of N-(3-dimetilaminopropil)-N′-etilcarbodiimide hydrochloride (EDC·HCl) in DMSO. After a 24 h reaction at 15 °C, the mixture was precipitated in diethyl ether/acetone (v/v = 1/2). The product was further purified using a dialysis tube (molecular weight cut-off, MWCO 3500 Da) with distilled water. The final product was obtained by lyophilization.

5. Conclusions

The main scope of this review was to attract more attention to a class of quaternary “onium” salts, such as quaternary phosphonium salts (QPSs), extensively studied for other purposes but still little considered as new antibacterial agents in comparison to their nitrogen-based counterpart. In this context, the study of QPSs as possible next-generation weapons to prevent BF formation or eradicate mature BFs is even more limited. In fact, we have underlined that in the last 15 years, publications concerning the anti-biofilm effects of QPSs total only 13 vs. the 602 concerning quaternary ammonium salts (QASs). However, the paucity of innovation in this space has driven the emergence of QAC resistance by different mechanisms including the overexpression of efflux pumps. On the contrary, there could be several advantages of employing QPSs in place of QASs, including less tendency to develop resistance, low toxicity towards eukaryotic cells, higher antibacterial and biocidal effects, capacity of quorum sensing (QS) disruption, and higher thermal stability. Additionally, the synthetic procedures useful to prepare the most promising antibacterial QPSs developed so far have been reviewed, to stimulate scientists to synthesize new QPS derivatives and enlarge the research concerning these compounds, which are still too little studied, and their antimicrobial effects. We are confident that the extension of knowledge on these materials by this review could be a successful approach to finding effective new weapons for treating chronic infections sustained by biofilm-producing MDR pathogens and environmental issues associated with biofilm-sustained fouling.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/pharmaceutics16010080/s1, Table S1: Clinically relevant MDR pathogens with the associated infections [7]; Table S2: General composition of BFs; Table S3: Stages of the BF life cycle.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available in this article (and Supplementary Materials).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cook, M.A.; Wright, G.D. The past, present, and future of antibiotics. Sci. Transl. Med. 2022, 14, eabo7793. [Google Scholar] [CrossRef] [PubMed]
  2. CDC. Center for Disease Control and Prevention. About Antimicrobial Resistance. Available online: https://www.cdc.gov/drugresistance/about.html (accessed on 23 November 2023).
  3. Chinemerem Nwobodo, D.; Ugwu, M.C.; Oliseloke Anie, C.; Al-Ouqaili, M.T.S.; Chinedu Ikem, J.; Victor Chigozie, U.; Saki, M. Antibiotic Resistance: The Challenges and Some Emerging Strategies for Tackling a Global Menace. J. Clin. Lab. Anal. 2022, 36, e24655. [Google Scholar] [CrossRef] [PubMed]
  4. Tang, K.W.K.; Millar, B.C.; Moore, J.E. Antimicrobial Resistance (AMR). Br. J. Biomed. Sci. 2023, 80, 11387. [Google Scholar] [CrossRef] [PubMed]
  5. Karakonstantis, S.; Kritsotakis, E.I.; Gikas, A. Treatment options for K. pneumoniae, P. aeruginosa and A. baumannii co-resistant to carbapenems, aminoglycosides, polymyxins and tigecycline: An approach based on the mechanisms of resistance to carbapenems. Infection 2020, 48, 835–851. [Google Scholar] [CrossRef] [PubMed]
  6. Munita, J.M.; Arias, C.A. Mechanisms of Antibiotic Resistance. Microbiol. Spectr. 2016, 4, 481–511. [Google Scholar] [CrossRef] [PubMed]
  7. Alfei, S.; Caviglia, D. Prevention and Eradication of Biofilm by Dendrimers: A Possibility Still Little Explored. Pharmaceutics 2022, 14, 2016. [Google Scholar] [CrossRef]
  8. Pang, Z.; Raudonis, R.; Glick, B.R.; Lin, T.-J.; Cheng, Z. Antibiotic resistance in Pseudomonas aeruginosa: Mechanisms and alternative therapeutic strategies. Biotechnol. Adv. 2019, 37, 177–192. [Google Scholar] [CrossRef]
  9. Jia, W.; Li, G.; Wang, W. Prevalence and Antimicrobial Resistance of Enterococcus Species: A Hospital-Based Study in China. Int. J. Environ. Res. Public Health 2014, 11, 3424–3442. [Google Scholar] [CrossRef]
  10. Grassotti, T.T.; de Angelis, Z.D.; da Fontoura Xavier Costa, L.; de Araújo, A.J.G.; Pereira, R.I.; Soares, R.O.; Wagner, P.G.C.; Frazzon, J.; Frazzon, A.P.G. Antimicrobial Resistance Profiles in Enterococcus spp. Isolates from Fecal Samples of Wild and Captive Black Capuchin Monkeys (Sapajus nigritus) in South Brazil. Front. Microbiol. 2018, 9, 2366. [Google Scholar] [CrossRef]
  11. Kyriakidis, I.; Vasileiou, E.; Pana, Z.D.; Tragiannidis, A. Acinetobacter baumannii Antibiotic Resistance Mechanisms. Pathogens 2021, 10, 373. [Google Scholar] [CrossRef]
  12. Galindo-Méndez, M. Antimicrobial Resistance in Escherichia coli. In E. Coli Infections-Importance of Early Diagnosis and Efficient Treatment; Louis, R., Ed.; IntechOpen: London, UK, 2020; p. 234. [Google Scholar] [CrossRef]
  13. Moya, C.; Maicas, S. Antimicrobial Resistance in Klebsiella pneumoniae Strains: Mechanisms and Outbreaks. Proceedings 2020, 66, 11. [Google Scholar] [CrossRef]
  14. Foster, T.J. Antibiotic Resistance in Staphylococcus aureus. Current Status and Future Prospects. FEMS Microbiol. Rev. 2017, 41, 430–449. [Google Scholar] [CrossRef] [PubMed]
  15. Cillóniz, C.; Garcia-Vidal, C.; Ceccato, A.; Torres, A. Antimicrobial Resistance Among Streptococcus pneumoniae. In Antimicrobial Resistance in the 21st Century; Fong, I.W., Shlaes, D., Drlica, K., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 13–38. ISBN 978-3-319-78538-7. [Google Scholar]
  16. Cuypers, W.L.; Jacobs, J.; Wong, V.; Klemm, E.J.; Deborggraeve, S.; Van Puyvelde, S. Fluoroquinolone Resistance in Salmonella: Insights by Whole-Genome Sequencing. Microb. Genom. 2018, 4, e000195. [Google Scholar] [CrossRef] [PubMed]
  17. Lampel, K.A. Chapter 7—Antimicrobial Resistance in Shigella Species. In Antimicrobial Resistance and Food Safety; Chen, C.-Y., Yan, X., Jackson, C.R., Eds.; Academic Press: San Diego, CA, USA, 2015; pp. 119–135. ISBN 978-0-12-801214-7. [Google Scholar]
  18. WHO. Multi-Drug Resistant Gonorrhoea. Available online: https://www.who.int/news-room/fact-sheets/detail/multi-drug-resistant-gonorrhoea (accessed on 21 November 2023).
  19. Li, H.; Yuan, J.; Duan, S.; Pang, Y. Resistance and Tolerance of Mycobacterium tuberculosis to Antimicrobial Agents—How M. Tuberculosis Can Escape Antibiotics. WIREs Mech. Dis. 2022, 14, e1573. [Google Scholar] [CrossRef] [PubMed]
  20. Loeffler, J.; Stevens, D.A. Antifungal Drug Resistance. Clin. Infect. Dis. 2003, 36, S31–S41. [Google Scholar] [CrossRef] [PubMed]
  21. Rodero, L.; Mellado, E.; Rodriguez, A.C.; Salve, A.; Guelfand, L.; Cahn, P.; Cuenca-Estrella, M.; Davel, G.; Rodriguez-Tudela, J.L. G484S Amino Acid Substitution in Lanosterol 14-α Demethylase (ERG11) Is Related to Fluconazole Resistance in a Recurrent Cryptococcus Neoformans Clinical Isolate. Antimicrob. Agents Chemother. 2003, 47, 3653–3656. [Google Scholar] [CrossRef] [PubMed]
  22. Howard, S.J.; Arendrup, M.C. Acquired Antifungal Drug Resistance in Aspergillus fumigatus: Epidemiology and Detection. Med. Mycol. 2011, 49, S90–S95. [Google Scholar] [CrossRef] [PubMed]
  23. Cuenca-Estrella, M.; Gomez-Lopez, A.; Mellado, E.; Buitrago, M.J.; Monzón, A.; Rodriguez-Tudela, J.L. Scopulariopsis brevicaulis, a Fungal Pathogen Resistant to Broad-Spectrum Antifungal Agents. Antimicrob. Agents Chemother. 2003, 47, 2339–2341. [Google Scholar] [CrossRef]
  24. Lurain, N.S.; Chou, S. Antiviral Drug Resistance of Human Cytomegalovirus. Clin. Microbiol. Rev. 2010, 23, 689–712. [Google Scholar] [CrossRef]
  25. Wutzler, P. Antiviral Therapy of Herpes Simplex and Varicella-Zoster Virus Infections. Intervirology 2008, 40, 343–356. [Google Scholar] [CrossRef]
  26. Cortez, K.J.; Maldarelli, F. Clinical Management of HIV Drug Resistance. Viruses 2011, 3, 347–378. [Google Scholar] [CrossRef]
  27. Hurt, A.C. The Epidemiology and Spread of Drug Resistant Human Influenza Viruses. Curr. Opin. Virol. 2014, 8, 22–29. [Google Scholar] [CrossRef]
  28. Suppiah, J.; Mohd Zain, R.; Haji Nawi, S.; Bahari, N.; Saat, Z. Drug-Resistance Associated Mutations in Polymerase (p) Gene of Hepatitis B Virus Isolated from Malaysian HBV Carriers. Hepat. Mon. 2014, 14, e13173. [Google Scholar] [CrossRef]
  29. Bloland, P.B. Drug Resistance in Malaria; A Background Document for the WHO Global Strategy for Containment of Antimicrobial Resistance; WHO: Geneva, Switzerland, 2001. [Google Scholar]
  30. Vanaerschot, M.; Dumetz, F.; Roy, S.; Ponte-Sucre, A.; Arevalo, J.; Dujardin, J.-C. Treatment Failure in Leishmaniasis: Drug-Resistance or Another (Epi-) Phenotype? Expert Rev. Anti-Infect. Ther. 2014, 12, 937–946. [Google Scholar] [CrossRef]
  31. Mohapatra, S. Drug Resistance in Leishmaniasis: Newer Developments. Trop. Parasitol. 2014, 4, 4–9. [Google Scholar] [CrossRef]
  32. Fallon, P.G.; Doenhoff, M.J. Drug-Resistant Schistosomiasis: Resistance to Praziquantel and Oxamniquine Induced in Schistosoma Mansoni in Mice Is Drug Specific. Am. J. Trop. Med. Hyg. 1994, 51, 83–88. [Google Scholar] [CrossRef]
  33. Qi, L.; Cui, J. A Schistosomiasis Model with Praziquantel Resistance. Discret. Dyn. Nat. Soc. 2013, 2013, 945767. [Google Scholar] [CrossRef]
  34. Bansal, D.; Malla, N.; Mahajan, R. Drug Resistance in Amoebiasis. Indian J. Med. Res. 2006, 123, 115–118. [Google Scholar]
  35. Muzny, C.A.; Schwebke, J.R. The Clinical Spectrum of Trichomonas Vaginalis Infection and Challenges to Management. Sex. Transm. Infect. 2013, 89, 423–425. [Google Scholar] [CrossRef]
  36. McFadden, D.C.; Tomavo, S.; Berry, E.A.; Boothroyd, J.C. Characterization of Cytochrome b from Toxoplasma Gondii and Qo Domain Mutations as a Mechanism of Atovaquone-Resistance. Mol. Biochem. Parasitol. 2000, 108, 1–12. [Google Scholar] [CrossRef]
  37. Nagamune, K.; Moreno, S.N.; Sibley, L.D. Artemisinin-Resistant Mutants of Toxoplasma gondii Have Altered Calcium Homeostasis. Antimicrob. Agents Chemother. 2007, 51, 3816–3823. [Google Scholar] [CrossRef]
  38. Doliwa, C.; Escotte-Binet, S.; Aubert, D.; Sauvage, V.; Velard, F.; Schmid, A.; Villena, I. Sulfadiazine Resistance in Toxoplasma gondii: No Involvement of Overexpression or Polymorphisms in Genes of Therapeutic Targets and ABC Transporters. Parasite 2013, 20, 19. [Google Scholar] [CrossRef]
  39. Merchel Piovesan Pereira, B.; Tagkopoulos, I. Benzalkonium Chlorides: Uses, Regulatory Status, and Microbial Resistance. Appl. Environ. Microbiol. 2019, 85, e00377-19. [Google Scholar] [CrossRef]
  40. Osmanov, A.; Farooq, Z.; Richardson, M.D.; Denning, D.W. The antiseptic Miramistin: A review of its comparative in vitro and clinical activity. FEMS Microbiol. Rev. 2020, 44, 399–417. [Google Scholar] [CrossRef]
  41. Kampf, G. Didecyldimethylammonium Chloride. In Antiseptic Stewardship: Biocide Resistance and Clinical Implications; Kampf, G., Ed.; Springer International Publishing: Cham, Switzerland, 2018; pp. 371–394. [Google Scholar]
  42. Mao, X.; Auer David, L.; Buchalla, W.; Hiller, K.-A.; Maisch, T.; Hellwig, E.; Al-Ahmad, A.; Cieplik, F. Cetylpyridinium Chloride: Mechanism of Action, Antimicrobial Efficacy in Biofilms, and Potential Risks of Resistance. Antimicrob. Agents Chemother. 2020, 64, e00576-20. [Google Scholar] [CrossRef]
  43. Assadian, O. Octenidine dihydrochloride: Chemical characteristics and antimicrobial properties. J. Wound Care 2016, 25, S3–S6. [Google Scholar] [CrossRef]
  44. Kwaśniewska, D.; Chen, Y.-L.; Wieczorek, D. Biological Activity of Quaternary Ammonium Salts and Their Derivatives. Pathogens 2020, 9, 459. [Google Scholar] [CrossRef]
  45. Michaud, M.E.; Allen, R.A.; Morrison-Lewis, K.R.; Sanchez, C.A.; Minbiole, K.P.C.; Post, S.J.; Wuest, W.M. Quaternary Phosphonium Compound Unveiled as a Potent Disinfectant against Highly Resistant Acinetobacter baumannii Clinical Isolates. ACS Infect. Dis. 2022, 8, 2307–2314. [Google Scholar] [CrossRef]
  46. 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]
  47. Santos, A.L.S.; dos Galdino, A.C.M.; Mello, T.P.; de Ramos, L.D.S.; Branquinha, M.H.; Bolognese, A.M.; Columbano Neto, J.; Roudbary, M. What Are the Advantages of Living in a Community? A Microbial Biofilm Perspective! Memórias Do Inst. Oswaldo Cruz 2018, 113, e180212. [Google Scholar] [CrossRef]
  48. Penesyan, A.; Paulsen, I.T.; Kjelleberg, S.; Gillings, M.R. Three Faces of Biofilms: A Microbial Lifestyle, a Nascent Multicellular Organism, and an Incubator for Diversity. NPJ Biofilms Microbiomes 2021, 7, 80. [Google Scholar] [CrossRef]
  49. Flemming, H.-C.; Wuertz, S. Bacteria and Archaea on Earth and Their Abundance in Biofilms. Nat. Rev. Microbiol. 2019, 17, 247–260. [Google Scholar] [CrossRef]
  50. Dincer, S.; Özdenefe, M.S.; Arkut, A. Bacterial Biofilms; IntechOpen: Rijeka, Yugoslavia, 2020; ISBN 978-1-78985-900-3. [Google Scholar]
  51. Kostakioti, M.; Hadjifrangiskou, M.; Hultgren, S.J. Bacterial Biofilms: Development, Dispersal, and Therapeutic Strategies in the Dawn of the Postantibiotic Era. Cold Spring Harb. Perspect. Med. 2013, 3, a010306. [Google Scholar] [CrossRef]
  52. de Vos, W.M. Microbial Biofilms and the Human Intestinal Microbiome. NPJ Biofilms Microbiomes 2015, 1, 15005. [Google Scholar] [CrossRef]
  53. Carthey, A.J.R.; Blumstein, D.T.; Gallagher, R.V.; Tetu, S.G.; Gillings, M.R. Conserving the Holobiont. Funct. Ecol. 2020, 34, 764–776. [Google Scholar] [CrossRef]
  54. Jiang, Y.; Geng, M.; Bai, L. Targeting Biofilms Therapy: Current Research Strategies and Development Hurdles. Microorganisms 2020, 8, 1222. [Google Scholar] [CrossRef]
  55. Penesyan, A.; Gillings, M.; Paulsen, I.T. Antibiotic Discovery: Combatting Bacterial Resistance in Cells and in Biofilm Communities. Molecules 2015, 20, 5286–5298. [Google Scholar] [CrossRef]
  56. Donlan, R.M.; Costerton, J.W. Biofilms: Survival Mechanisms of Clinically Relevant Microorganisms. Clin. Microbiol. Rev. 2002, 15, 167–193. [Google Scholar] [CrossRef]
  57. Donlan, R.M. Biofilms and Device-Associated Infections. Emerg. Infect. Dis. 2001, 7, 277–281. [Google Scholar] [CrossRef]
  58. Percival, S.L.; Kite, P. Intravascular Catheters and Biofilm Control. J. Vasc. Access 2007, 8, 69–80. [Google Scholar] [CrossRef]
  59. Safdar, N.; Mermel, L.A.; Maki, D.G. The epidemiology of catheter-related infection in the critically ill. In Catheter-Related Infections in the Critically Ill; O’Grady, N., Pittet, D., Eds.; Kluwer: New York, NY, USA, 2004; pp. 1–23. [Google Scholar]
  60. Donlan, R.; Murga, R.; Carson, L. Growing biofilms in intravenous fluids. In Biofilms, the Good, the Bad, and the Ugly; Wimpenny, J., Gilbert, P., Walker, J., Brading, M., Bayston, R., Eds.; Biofilm Club: Powys, UK, 1999; pp. 23–29. [Google Scholar]
  61. Chow, K.M.; Li, P.K.-T.; Cho, Y.; Abu-Alfa, A.; Bavanandan, S.; Brown, E.A.; Cullis, B.; Edwards, D.; Ethier, I.; Hurst, H.; et al. ISPD Catheter-Related Infection Recommendations: 2023 Update. Perit. Dial. Int. 2023, 43, 201–219. [Google Scholar] [CrossRef] [PubMed]
  62. Zheng, J.M.D.; Romanelli, A.M. An Overview of Prosthetic Joint Infection (PJI) Definition and Diagnosis. 13 July 2021. Available online: https://health.ucdavis.edu/blog/lab-best-practice/an-overview-of-prosthetic-joint-infection-pji-definition-and-diagnosis/2021/07#:~:text=Pathogens%20involved%20in%20Prosthetic%20joint%20infection%201%20Early,after%20surgery%29%20Staphylococcus%20aureus%20Gram-negative%20bacilli%20Beta-hemolytic%20streptococci (accessed on 21 November 2023).
  63. Palmeri, N.O.; Kramer, D.B.; Karchmer, A.W.; Zimetbaum, P.J. A Review of Cardiac Implantable Electronic Device Infections for the Practicing Electrophysiologist. JACC Clin. Electrophysiol. 2021, 7, 811–824. [Google Scholar] [CrossRef] [PubMed]
  64. Kokare, C.R.; Chakraborty, S.; Khopade, A.; Mahadik, K.R. Biofilm: Importance and Applications. Indian J. Biotechnol. 2009, 8, 159–168. [Google Scholar]
  65. Stickler, D.J. Bacterial Biofilms and the Encrustation of Urethral Catheters. Biofouling 1996, 9, 293–305. [Google Scholar] [CrossRef]
  66. Pentland, D.R.; Stevens, S.; Williams, L.; Baker, M.; McCall, C.; Makarovaite, V.; Balfour, A.; Mühlschlegel, F.A.; Gourlay, C.W. Precision Antifungal Treatment Significantly Extends Voice Prosthesis Lifespan in Patients Following Total Laryngectomy. Front. Microbiol. 2020, 11, 975. [Google Scholar] [CrossRef] [PubMed]
  67. Kokare, C.R.; Kadam, S.S.; Mahadik, K.R.; Chopade, B.A. Studies on bioemulsifier production from marine Streptomyces sp. S1. Ind. J. Biotechnol. 2007, 6, 78–84. [Google Scholar]
  68. Lamont, R.J.; Jenkinson, H.F. Life Below the Gum Line: Pathogenic Mechanisms of Porphyromonas gingivalis. Microbiol. Mol. Biol. Rev. 1998, 62, 1244–1263. [Google Scholar] [CrossRef] [PubMed]
  69. Overman, P.R. Biofilm: A New View of Plaque. J. Contemp. Dent. Pract. 2000, 1, 18–29. [Google Scholar] [CrossRef]
  70. Nasser, A.; Azimi, T.; Ostadmohammadi, S.; Ostadmohammadi, S. A comprehensive review of bacterial osteomyelitis with emphasis on Staphylococcus aureus. Microb. Pathog. 2020, 148, 104431. [Google Scholar] [CrossRef]
  71. Mirzaei, R.; Mohammadzadeh, R.; Alikhani, M.Y.; Shokri Moghadam, M.; Karampoor, S.; Kazemi, S.; Barfipoursalar, A.; Yousefimashouf, R. The Biofilm-Associated Bacterial Infections Unrelated to Indwelling Devices. IUBMB Life 2020, 72, 1271–1285. [Google Scholar] [CrossRef]
  72. Arunasri, K.; Mohan, S.V. Chapter 2.3—Biofilms: Microbial Life on the Electrode Surface. In Microbial Electrochemical Technology; Mohan, S.V., Varjani, S., Pandey, A., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 295–313. ISBN 978-0-444-64052-9. [Google Scholar]
  73. Jamal, M.; Tasneem, U.; Hussain, T.; Andleeb, S. Bacterial Biofilm: Its Composition, Formation and Role in Human Infections. Res. Rev. J. Microbiol. Biotechnol. 2015, 4, 59. [Google Scholar]
  74. Monroe, D. Looking for Chinks in the Armor of Bacterial Biofilms. PLoS Biol. 2007, 5, e307. [Google Scholar] [CrossRef] [PubMed]
  75. Kaplan, J.B. Biofilm Dispersal: Mechanisms, Clinical Implications, and Potential Therapeutic Uses. J. Dent. Res. 2010, 89, 205–218. [Google Scholar] [CrossRef] [PubMed]
  76. Davies, D.G.; Marques, C.N.H. A Fatty Acid Messenger Is Responsible for Inducing Dispersion in Microbial Biofilms. J. Bacteriol. 2009, 191, 1393–1403. [Google Scholar] [CrossRef] [PubMed]
  77. Rice, S.A.; Koh, K.S.; Queck, S.Y.; Labbate, M.; Lam, K.W.; Kjelleberg, S. Biofilm Formation and Sloughing in Serratia marcescens Are Controlled by Quorum Sensing and Nutrient Cues. J. Bacteriol. 2005, 187, 3477–3485. [Google Scholar] [CrossRef] [PubMed]
  78. Preda, V.; Sandulescu, O. Communication Is the Key: Biofilms, Quorum Sensing, Formation and Prevention. Discoveries 2019, 7, e10. [Google Scholar] [CrossRef] [PubMed]
  79. Verma, S.C.; Miyashiro, T. Quorum Sensing in the Squid-Vibrio Symbiosis. Int. J. Mol. Sci. 2013, 14, 16386–16401. [Google Scholar] [CrossRef]
  80. Brackman, G.; Coenye, T. Quorum sensing inhibitors as anti-biofilm agents. Curr. Pharm. Des. 2015, 21, 5–11. [Google Scholar] [CrossRef]
  81. Li, Z.; Nair, S.K. Quorum Sensing: How Bacteria Can Coordinate Activity and Synchronize Their Response to External Signals? Protein Sci. 2012, 21, 1403–1417. [Google Scholar] [CrossRef]
  82. Jiang, Q.; Chen, J.; Yang, C.; Yin, Y.; Yao, K. Quorum Sensing: A Prospective Therapeutic Target for Bacterial Diseases. BioMed Res. Int. 2019, 2019, 2015978. [Google Scholar] [CrossRef]
  83. Schirmer, T. C-Di-GMP Synthesis: Structural Aspects of Evolution, Catalysis and Regulation. J. Mol. Biol. 2016, 428, 3683–3701. [Google Scholar] [CrossRef] [PubMed]
  84. Lin, Y.; Mi, D.; Hou, Y.; Lin, M.; Xie, Q.; Niu, X.; Chen, Y.; He, C.; Tao, J.; Li, C. Systematic Analysis of the Roles of C-Di-GMP Signaling in Xanthomonas oryzae Pv. Oryzae Virulence. FEMS Microbiol. Lett. 2022, 369, fnac068. [Google Scholar] [CrossRef] [PubMed]
  85. Valentini, M.; Filloux, A. Biofilms and Cyclic Di-GMP (c-Di-GMP) Signaling: Lessons from Pseudomonas aeruginosa and Other Bacteria. J. Biol. Chem. 2016, 291, 12547–12555. [Google Scholar] [CrossRef] [PubMed]
  86. Dieltjens, L.; Appermans, K.; Lissens, M.; Lories, B.; Kim, W.; Van der Eycken, E.V.; Foster, K.R.; Steenackers, H.P. Inhibiting Bacterial Cooperation Is an Evolutionarily Robust Anti-Biofilm Strategy. Nat. Commun. 2020, 11, 107. [Google Scholar] [CrossRef] [PubMed]
  87. Khatoon, Z.; McTiernan, C.D.; Suuronen, E.J.; Mah, T.-F.; Alarcon, E.I. Bacterial Biofilm Formation on Implantable Devices and Approaches to Its Treatment and Prevention. Heliyon 2018, 4, e01067. [Google Scholar] [CrossRef] [PubMed]
  88. Alfei, S.; Schito, A.M. From Nanobiotechnology, Positively Charged Biomimetic Dendrimers as Novel Antibacterial Agents: A Review. Nanomaterials 2020, 10, 2022. [Google Scholar] [CrossRef]
  89. Terreni, M.; Taccani, M.; Pregnolato, M. New Antibiotics for Multidrug-Resistant Bacterial Strains: Latest Research Developments and Future Perspectives. Molecules 2021, 26, 2671. [Google Scholar] [CrossRef]
  90. Mataraci, E.; Dosler, S. In Vitro Activities of Antibiotics and Antimicrobial Cationic Peptides Alone and in Combination against Methicillin-Resistant Staphylococcus aureus Biofilms. Antimicrob. Agents Chemother. 2012, 56, 6366–6371. [Google Scholar] [CrossRef]
  91. Dosler, S.; Mataraci, E. In Vitro Pharmacokinetics of Antimicrobial Cationic Peptides Alone and in Combination with Antibiotics against Methicillin Resistant Staphylococcus aureus Biofilms. Peptides 2013, 49, 53–58. [Google Scholar] [CrossRef]
  92. Dosler, S.; Karaaslan, E. Inhibition and Destruction of Pseudomonas aeruginosa Biofilms by Antibiotics and Antimicrobial Peptides. Peptides 2014, 62, 32–37. [Google Scholar] [CrossRef]
  93. Field, D.; O’ Connor, R.; Cotter, P.D.; Ross, R.P.; Hill, C. In Vitro Activities of Nisin and Nisin Derivatives Alone and In Combination with Antibiotics against Staphylococcus biofilms. Front. Microbiol. 2016, 7, 508. [Google Scholar] [CrossRef]
  94. Obłąk, E.; Gamian, A. Biologiczna aktywność czwartorzędowych soli amoniowych (CSA) * The biological activity of quaternary ammonium salts (QASs). Postepy Hig. Med. Dosw. 2010, 64, 201–211. [Google Scholar]
  95. Gerba, C.P. Quaternary ammonium biocides: Efficacy in application. Appl. Environ. Microbiol. 2015, 81, 464–469. [Google Scholar] [CrossRef] [PubMed]
  96. Cole, E.C.; Addison, R.M.; Rubino, J.R.; Leese, K.E.; Dulaney, P.D.; Newell, M.S.; Wilkins, J.; Gaber, D.J.; Wineinger, T.; Criger, D.A. Investigation of antibiotic and antibacterial agent cross-resistance in target bacteria from homes of antibacterial product users and nonusers. J. Appl. Microbiol. 2003, 95, 664–676. [Google Scholar] [CrossRef] [PubMed]
  97. Mbithi, J.N.; Springthorpe, V.S.; Sattar, S.A. Chemical disinfection of hepatitis A virus on environmental surfaces. Appl. Environ. Microbiol. 1990, 56, 3601–3604. [Google Scholar] [CrossRef] [PubMed]
  98. Tischer, M.; Pradel, G.; Ohlsen, K.; Holzgrabe, U. Quaternary ammonium salts and their antimicrobial potential: Targets or nonspecific interactions? Chem. Med. Chem. 2012, 7, 22–31. [Google Scholar] [CrossRef] [PubMed]
  99. Gilbert, P.; Moore, L.E. Cationic antiseptics: Diversity of action under a common epithet. J. Appl. Microbiol. 2005, 99, 703–715. [Google Scholar] [CrossRef]
  100. Buffet-Bataillon, S.; Tattevin, P.; Bonnaure-Mallet, M.; Jolivet-Gougeon, A. Emergence of resistance to antibacterial agents: The role of quaternary ammonium compounds—A critical review. Int. J. Antimicrob. Agents 2012, 39, 381–389. [Google Scholar] [CrossRef]
  101. Bragg, R.; Jansen, A.; Coetzee, M.; van der Westhuizen, W.; Boucher, C. Bacterial resistance to quaternary ammonium compounds (QAC) disinfectants. In Infectious Diseases and Nanomedicine II; Adhikari, R., Thapa, S., Eds.; Springer: New Delhi, India, 2014; Volume 808, pp. 1–13. [Google Scholar]
  102. Peng, Z.-X.; Tu, B.; Shen, Y.; Du, L.; Wang, L.; Guo, S.-R.; Tang, T.-T. Quaternized Chitosan Inhibits icaA Transcription and Biofilm Formation by Staphylococcus on a Titanium Surface. Antimicrob. Agents Chemother. 2011, 55, 860–866. [Google Scholar] [CrossRef]
  103. Tan, H.; Peng, Z.; Li, Q.; Xu, X.; Guo, S.; Tang, T. The Use of Quaternised Chitosan-Loaded PMMA to Inhibit Biofilm Formation and Downregulate the Virulence-Associated Gene Expression of Antibiotic-Resistant Staphylococcus. Biomaterials 2012, 33, 365–377. [Google Scholar] [CrossRef]
  104. Song, J.; Kong, H.; Jang, J. Bacterial Adhesion Inhibition of the Quaternary Ammonium Functionalized Silica Nanoparticles. Colloids Surf. B Biointerfaces 2011, 82, 651–656. [Google Scholar] [CrossRef] [PubMed]
  105. Obłąk, E.; Piecuch, A.; Rewak-Soroczyńska, J.; Paluch, E. Activity of gemini quaternary ammonium salts against microorganisms. Appl. Microbiol. Biotechnol. 2019, 103, 625–632. [Google Scholar] [CrossRef] [PubMed]
  106. Obłąk, E.; Piecuch, A.; Krasowska, A.; Łuczyński, J. Antifungal activity of gemini quaternary ammonium salts. Microbiol. Res. 2013, 168, 630–638. [Google Scholar] [CrossRef] [PubMed]
  107. Tran, P.L.; Hamood, A.N.; de Souza, A.; Schultz, G.; Liesenfeld, B.; Mehta, D.; Reid, T.W. A study on the ability of quaternary ammonium groups attached to a polyurethane foam wound dressing to inhibit bacterial attachment and biofilm formation. Wound Repair Regen. 2015, 23, 74–81. [Google Scholar] [CrossRef] [PubMed]
  108. Kocak, H.S.; Bulut, O.; Yilmaz, M.D. A Dicationic BODIPY-Based Fluorescent Bactericide to Combat Infectious Diseases and to Eradicate Bacterial Biofilms. ACS Appl. BioMaterials 2023, 6, 1604–1610. [Google Scholar] [CrossRef]
  109. Ibrahim, M.S.; Ibrahim, A.S.; Balhaddad, A.A.; Weir, M.D.; Lin, N.J.; Tay, F.R.; Oates, T.W.; Xu, H.H.K.; Melo, M.A.S. A Novel Dental Sealant Containing Dimethylaminohexadecyl Methacrylate Suppresses the Cariogenic Pathogenicity of Streptococcus Mutans Biofilms. Int. J. Mol. Sci. 2019, 20, 3491. [Google Scholar] [CrossRef] [PubMed]
  110. Daood, U.; Matinlinna, J.P.; Pichika, M.R.; Mak, K.-K.; Nagendrababu, V.; Fawzy, A.S. A Quaternary Ammonium Silane Antimicrobial Triggers Bacterial Membrane and Biofilm Destruction. Sci. Rep. 2020, 10, 10970. [Google Scholar] [CrossRef]
  111. Balbinot, G.S.; Marcon, N.; Sauro, S.; Luxan, S.A.; Collares, F.M. Alkyl Trimethyl Ammonium Bromide for the Formulation of Antibacterial Orthodontic Resins. Clin. Oral Investig. 2022, 26, 7011–7019. [Google Scholar] [CrossRef]
  112. Phuangkaew, T.; Booranabunyat, N.; Kiatkamjornwong, S.; Thanyasrisung, P.; Hoven, V.P. Amphiphilic Quaternized Chitosan: Synthesis, Characterization, and Anti-Cariogenic Biofilm Property. Carbohydr. Polym. 2022, 277, 118882. [Google Scholar] [CrossRef]
  113. He, Y.; Wan, X.; Xiao, K.; Lin, W.; Li, J.; Li, Z.; Luo, F.; Tan, H.; Li, J.; Fu, Q. Anti-Biofilm Surfaces from Mixed Dopamine-Modified Polymer Brushes: Synergistic Role of Cationic and Zwitterionic Chains to Resist: Staphyloccocus aureus. Biomater. Sci. 2019, 7, 5369–5382. [Google Scholar] [CrossRef]
  114. Yu, J.; Huang, X.; Zhou, X.; Han, Q.; Zhou, W.; Liang, J.; Xu, H.H.K.; Ren, B.; Peng, X.; Weir, M.D.; et al. Anti-Caries Effect of Resin Infiltrant Modified by Quaternary Ammonium Monomers. J. Dent. 2020, 97, 103355. [Google Scholar] [CrossRef] [PubMed]
  115. Fedorowicz, J.; Cruz, C.D.; Morawska, M.; Ciura, K.; Gilbert-Girard, S.; Mazur, L.; Mäkkylä, H.; Ilina, P.; Savijoki, K.; Fallarero, A.; et al. Antibacterial and Antibiofilm Activity of Permanently Ionized Quaternary Ammonium Fluoroquinolones. European J. Med. Chem. 2023, 254, 115373. [Google Scholar] [CrossRef] [PubMed]
  116. Daood, U.; Bapat, R.A.; Sidhu, P.; Ilyas, M.S.; Khan, A.S.; Mak, K.-K.; Pichika, M.R.; Nagendrababu, V.; Peters, O.A. Antibacterial and Antibiofilm Efficacy of K21-E in Root Canal Disinfection. Dent. Mater. 2021, 37, 1511–1528. [Google Scholar] [CrossRef] [PubMed]
  117. Piras, A.M.; Esin, S.; Benedetti, A.; Maisetta, G.; Fabiano, A.; Zambito, Y.; Batoni, G. Antibacterial, Antibiofilm, and Antiadhesive Properties of Different Quaternized Chitosan Derivatives. Int. J. Mol. Sci. 2019, 20, 6297. [Google Scholar] [CrossRef] [PubMed]
  118. Fugolin, A.P.; Dobson, A.; Huynh, V.; Mbiya, W.; Navarro, O.; Franca, C.M.; Logan, M.; Merritt, J.L.; Ferracane, J.L.; Pfeifer, C.S. Antibacterial, Ester-Free Monomers: Polymerization Kinetics, Mechanical Properties, Biocompatibility and Anti-Biofilm Activity. Acta Biomater. 2019, 100, 132–141. [Google Scholar] [CrossRef] [PubMed]
  119. Rita Pereira, A.; Gomes, I.B.; Simões, M. Choline-Based Ionic Liquids for Planktonic and Biofilm Growth Control of Bacillus cereus and Pseudomonas fluorescens. J. Mol. Liq. 2022, 346, 117077. [Google Scholar] [CrossRef]
  120. Gaspar, C.; Rolo, J.; Cerca, N.; Palmeira de Oliveira, R.; Martinez-de-Oliveira, J.; Palmeira-de-Oliveira, A. Dequalinium Chloride Effectively Disrupts Bacterial Vaginosis (BV) Gardnerella Spp. Biofilms. Pathogens 2021, 10, 261. [Google Scholar] [CrossRef]
  121. Li, Y.; Wang, H.; Zheng, X.; Li, Z.; Wang, M.; Luo, K.; Zhang, C.; Xia, X.; Wang, Y.; Shi, C. Didecyldimethylammonium Bromide: Application to Control Biofilms of Staphylococcus aureus and Pseudomonas aeruginosa Alone and in Combination with Slightly Acidic Electrolyzed Water. Food Res. Int. 2022, 157, 111236. [Google Scholar] [CrossRef]
  122. Yu, W.; Ren, C.; Zhang, N.; Cao, L.; Weir, M.D.; Yang, K.; Xu, H.H.K.; Bai, Y. Dual Function of Anti-Biofilm and Modulating Biofilm Equilibrium of Orthodontic Cement Containing Quaternary Ammonium Salt. Dent. Mater. J. 2023, 42, 149–157. [Google Scholar] [CrossRef]
  123. Moshynets, O.V.; Baranovskyi, T.P.; Iungin, O.S.; Kysil, N.P.; Metelytsia, L.O.; Pokholenko, I.; Potochilova, V.V.; Potters, G.; Rudnieva, K.L.; Rymar, S.Y.; et al. eDNA Inactivation and Biofilm Inhibition by the Polymeric Biocide Polyhexamethylene Guanidine Hydrochloride (PHMG-Cl). Int. J. Mol. Sci. 2022, 23, 731. [Google Scholar] [CrossRef]
  124. Daood, U.; Burrow, M.F.; Yiu, C.K.Y. Effect of a Novel Quaternary Ammonium Silane Cavity Disinfectant on Cariogenic Biofilm Formation. Clin. Oral Investig. 2020, 24, 649–661. [Google Scholar] [CrossRef] [PubMed]
  125. Feng, J.; Cheng, L.; Zhou, X.; Xu, H.H.K.; Weir, M.D.; Li, Q.; Hannig, M.; Rupf, S. Effects of Water Aging on the Mechanical and Anti-Biofilm Properties of Glass-Ionomer Cement Containing Dimethylaminododecyl Methacrylate. Dent. Mater. 2019, 35, 434–443. [Google Scholar] [CrossRef] [PubMed]
  126. Zhou, C.; Song, H.; Loh, J.L.C.; She, J.; Deng, L.; Liu, B. Grafting Antibiofilm Polymer Hydrogel Film onto Catheter by SARA SI-ATRP. J. Biomater. Sci. Polym. Ed. 2018, 29, 2106–2123. [Google Scholar] [CrossRef] [PubMed]
  127. Kuppusamy, R.; Yasir, M.; Yee, E.; Willcox, M.; Black, D.S.C.; Kumar, N. Guanidine Functionalized Anthranilamides as Effective Antibacterials with Biofilm Disruption Activity. Org. Biomol. Chem. 2018, 16, 5871–5888. [Google Scholar] [CrossRef] [PubMed]
  128. Melo, M.A.S.; Weir, M.D.; Passos, V.F.; Rolim, J.P.M.; Lynch, C.D.; Rodrigues, L.K.A.; Xu, H.H.K. Human in Situ Study of the Effect of Bis(2-Methacryloyloxyethyl) Dimethylammonium Bromide Immobilized in Dental Composite on Controlling Mature Cariogenic Biofilm. Int. J. Mol. Sci. 2018, 19, 3443. [Google Scholar] [CrossRef] [PubMed]
  129. Frolov, N.; Detusheva, E.; Fursova, N.; Ostashevskaya, I.; Vereshchagin, A. Microbiological Evaluation of Novel Bis-Quaternary Ammonium Compounds: Clinical Strains, Biofilms, and Resistance Study. Pharmaceuticals 2022, 15, 514. [Google Scholar] [CrossRef]
  130. Soares, N.S.; Hollanda, L.M.; Elias, C.M.V.; Azevedo, M.M.F.; Santos, F.E.P.; Lobo, A.O.; Marciano, F.R. Modified PCL/PEG/GelMA Electrospun Blends Reduced Biofilm Formation. Mater. Lett. 2022, 320, 132315. [Google Scholar] [CrossRef]
  131. Ran, H.-H.; Cheng, X.; Bao, Y.-W.; Hua, X.-W.; Gao, G.; Zhang, X.; Jiang, Y.-W.; Zhu, Y.-X.; Wu, F.-G. Multifunctional Quaternized Carbon Dots with Enhanced Biofilm Penetration and Eradication Efficiencies. J. Mater. Chem. B 2019, 7, 5104–5114. [Google Scholar] [CrossRef]
  132. Li, Y.; Zhen, J.; Tian, Q.; Shen, C.; Zhang, L.; Yang, K.; Shang, L. One Step Synthesis of Positively Charged Gold Nanoclusters as Effective Antimicrobial Nanoagents against Multidrug-Resistant Bacteria and Biofilms. J. Colloid Interface Sci. 2020, 569, 235–243. [Google Scholar] [CrossRef]
  133. Yang, J.; Zhu, Y.-X.; Lu, P.; Zhu, B.; Wu, F.-G. One-Step Synthesis of Quaternized Silica Nanoparticles with Bacterial Adhesion and Aggregation Properties for Effective Antibacterial and Antibiofilm Treatments. J. Mater. Chem. B 2022, 10, 3073–3082. [Google Scholar] [CrossRef]
  134. Xia, L.; Tian, J.; Yue, T.; Cao, H.; Chu, J.; Cai, H.; Zhang, W. Pillar [5]Arene-Based Acid-Triggered Supramolecular Porphyrin Photosensitizer for Combating Bacterial Infections and Biofilm Dispersion. Adv. Healthc. Mater. 2022, 11, 2102015. [Google Scholar] [CrossRef] [PubMed]
  135. Lan, T.; Guo, Q.; Shen, X. Polyethyleneimine and Quaternized Ammonium Polyethyleneimine: The Versatile Materials for Combating Bacteria and Biofilms. J. Biomater. Sci. Polym. Ed. 2019, 30, 1243–1259. [Google Scholar] [CrossRef] [PubMed]
  136. Yu, T.T.; Kuppusamy, R.; Yasir, M.; Hassan, M.M.; Sara, M.; Ho, J.; Willcox, M.D.P.; Black, D.S.; Kumar, N. Polyphenylglyoxamide-based Amphiphilic Small Molecular Peptidomimetics as Antibacterial Agents with Anti-biofilm Activity. Int. J. Mol. Sci. 2021, 22, 7344. [Google Scholar] [CrossRef] [PubMed]
  137. Balhaddad, A.A.; Ibrahim, M.S.; Garcia, I.M.; Collares, F.M.; Weir, M.D.; Xu, H.H.; Melo, M.A.S. Pronounced Effect of Antibacterial Bioactive Dental Composite on Microcosm Biofilms Derived from Patients with Root Carious Lesions. Front. Mater. 2020, 7, 583861. [Google Scholar] [CrossRef]
  138. He, D.; Yu, Y.; Liu, F.; Yao, Y.; Li, P.; Chen, J.; Ning, N.; Zhang, S. Quaternary Ammonium Salt-Based Cross-Linked Micelle Templated Synthesis of Highly Active Silver Nanocomposite for Synergistic Anti-Biofilm Application. Chem. Eng. J. 2020, 382, 122976. [Google Scholar] [CrossRef]
  139. Wu, G.-Y.; Yu, L.; Wang, Y.-R.; Yuan, X.; Tang, Y.-F.; Chen, W.; Zeng, L.-Z. Quaternary Ammonium Salt-Based Cross-Linked Micelle with Copper Nanoparticles for Treatment of Sulfate Reducing Bacteria Biofilm. React. Funct. Polym. 2022, 180, 105405. [Google Scholar] [CrossRef]
  140. Zhou, Y.; Wang, S.; Zhou, X.; Zou, Y.; Li, M.; Peng, X.; Ren, B.; Xu, H.H.K.; Weir, M.D.; Cheng, L.; et al. Short-Time Antibacterial Effects of Dimethylaminododecyl Methacrylate on Oral Multispecies Biofilm in Vitro. BioMed Res. Int. 2019, 2019, 6393470. [Google Scholar] [CrossRef]
  141. Hernando-Gozalo, M.; Aguilera-Correa, J.J.; Rescalvo-Casas, C.; Seijas-Pereda, L.; García-Bertolín, C.; de la Mata, F.J.; Sánchez-Nieves, J.; Cuadros, J.; Pérez-Tanoira, R. Study of the Antimicrobial Activity of Cationic Carbosilane Dendrimers against Clinical Strains of Multidrug-Resistant Bacteria and Their Biofilms. Front. Cell. Infect. Microbiol. 2023, 13, 3991. [Google Scholar] [CrossRef]
  142. Gao, S.; Sun, Y.; Lu, Z.; Jiang, N.; Yao, H. Synergistic Antibacterial and Biofilm Eradication Activity of Quaternary-Ammonium Compound with Copper Ion. J. Inorg. Biochem. 2023, 243, 112190. [Google Scholar] [CrossRef]
  143. Ali, I.; Burki, S.; El-Haj, B.M.; Parveen, S.; Nadeem, H.Ş.; Nadeem, S.; Shah, M.R. Synthesis and Characterization of Pyridine-Based Organic Salts: Their Antibacterial, Antibiofilm and Wound Healing Activities. Bioorganic Chem. 2020, 100, 103937. [Google Scholar] [CrossRef]
  144. Garipov, M.R.; Sabirova, A.E.; Pavelyev, R.S.; Shtyrlin, N.V.; Lisovskaya, S.A.; Bondar, O.V.; Laikov, A.V.; Romanova, J.G.; Bogachev, M.I.; Kayumov, A.R.; et al. Targeting Pathogenic Fungi, Bacteria and Fungal-Bacterial Biofilms by Newly Synthesized Quaternary Ammonium Derivative of Pyridoxine and Terbinafine with Dual Action Profile. Bioorganic Chem. 2020, 104, 104306. [Google Scholar] [CrossRef] [PubMed]
  145. Hympanova, M.; Terlep, S.; Markova, A.; Prchal, L.; Dogsa, I.; Pulkrabkova, L.; Benkova, M.; Marek, J.; Stopar, D. The Antibacterial Effects of New N-Alkylpyridinium Salts on Planktonic and Biofilm Bacteria. Front. Microbiol. 2020, 11, 573951. [Google Scholar] [CrossRef] [PubMed]
  146. Frolov, N.A.; Seferyan, M.A.; Valeev, A.B.; Saverina, E.A.; Detusheva, E.V.; Vereshchagin, A.N. The Antimicrobial and Antibiofilm Potential of New Water-Soluble Tris-Quaternary Ammonium Compounds. Int. J. Mol. Sci. 2023, 24, 10512. [Google Scholar] [CrossRef] [PubMed]
  147. Koziróg, A.; Otlewska, A.; Brycki, B. Viability, Enzymatic and Protein Profiles of Pseudomonas aeruginosa Biofilm and Planktonic Cells after Monomeric/Gemini Surfactant Treatment. Molecules 2018, 23, 1294. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, L.; Xin, M.; Li, M.; Liu, W.; Mao, Y. Effect of the Structure of Chitosan Quaternary Phosphonium Salt and Chitosan Quaternary Ammonium Salt on the Antibacterial and Antibiofilm Activity. Int. J. Biol. Macromol. 2023, 242, 124877. [Google Scholar] [CrossRef]
  149. Keasler, V.; Bennett, B.; McGinley, H. Analysis of Bacterial Kill Versus Corrosion from Use of Common Oilfield Biocides. In Proceedings of the 2010 8th International Pipeline Conference, Calgary, AB, Canada, 27 September–1 October 2010; ASME: New York, NY, USA; Volume 1, pp. 935–944. [Google Scholar] [CrossRef]
  150. Kramer, J.F.; Arpita, S.; Strba, S.F. Comparative Performance of Biocides Versus Corrosion Causing Biofilms. In NACE CORROSION; NACE: San Antonio, TX, USA, 2010. [Google Scholar]
  151. Kayumov, A.R.; Nureeva, A.A.; Trizna, E.Y.; Gazizova, G.R.; Bogachev, M.I.; Shtyrlin, N.V.; Pugachev, M.V.; Sapozhnikov, S.V.; Shtyrlin, Y.G. New Derivatives of Pyridoxine Exhibit High Antibacterial Activity against Biofilm-Embedded Staphylococcus Cells. BioMed Res. Int. 2015, 2015, 890968. [Google Scholar] [CrossRef]
  152. Joseph, R.; Naugolny, A.; Feldman, M.; Herzog, I.M.; Fridman, M.; Cohen, Y. Cationic Pillararenes Potently Inhibit Biofilm Formation without Affecting Bacterial Growth and Viability. J. Am. Chem. Soc. 2016, 138, 754–757. [Google Scholar] [CrossRef]
  153. Joseph, R.; Kaizerman, D.; Herzog, I.M.; Hadar, M.; Feldman, M.; Fridman, M.; Cohen, Y. Phosphonium Pillar [5]Arenes as a New Class of Efficient Biofilm Inhibitors: Importance of Charge Cooperativity and the Pillar Platform. Chem. Commun. 2016, 52, 10656–10659. [Google Scholar] [CrossRef]
  154. Das, S.; Paul, A.; Bera, D.; Dey, A.; Roy, A.; Dutta, A.; Ganguly, D. Design, Development and Mechanistic Insights into the Enhanced Antibacterial Activity of Mono and Bis-Phosphonium Fluoresceinate Ionic Liquids. Mater. Today Commun. 2021, 28, 102672. [Google Scholar] [CrossRef]
  155. Cieniecka-Rosłonkiewicz, A.; Pernak, J.; Kubis-Feder, J.; Ramani, A.; Robertson, A.J.; Seddon, K.R. Synthesis, Anti-Microbial Activities and Anti-Electrostatic Properties of Phosphonium-Based Ionic Liquids. Green Chem. 2005, 7, 855–862. [Google Scholar] [CrossRef]
  156. Shive, P.N.; Diehl, J.F. Reduction of Hematite to Magnetite under Natural and Laboratory Conditions. J. Geomagn. Geoelectr. 1977, 29, 345–354. [Google Scholar] [CrossRef]
  157. Atefi, F.; Garcia, M.T.; Singer, R.D.; Scammells, P.J. Phosphonium Ionic Liquids: Design, Synthesis and Evaluation of Biodegradability. Green Chem. 2009, 11, 1595–1604. [Google Scholar] [CrossRef]
  158. Metelytsia, L.O.; Hodyna, D.M.; Semenyuta, I.V.; Kovalishyn, V.V.; Rogalsky, S.P.; Derevianko, K.Y.; Brovarets, V.S.; Tetko, I.V. Theoretical and Experimental Studies of Phosphonium Ionic Liquids as Potential Antibacterials of MDR Acinetobacter baumannii. Antibiotics 2022, 11, 491. [Google Scholar] [CrossRef] [PubMed]
  159. Hemp, S.T.; Zhang, M.; Allen, M.H.; Cheng, S.; Moore, R.B.; Long, T.E. Comparing Ammonium and Phosphonium Polymerized Ionic Liquids: Thermal Analysis, Conductivity, and Morphology. Macromol. Chem. Phys. 2013, 214, 2099–2107. [Google Scholar] [CrossRef]
  160. Kanazawa, A.; Ikeda, T.; Endo, T. Synthesis and Antimicrobial Activity of Dimethyl- and Trimethyl-Substituted Phosphonium Salts with Alkyl Chains of Various Lengths. Antimicrob. Agents Chemother. 1994, 38, 945–952. [Google Scholar] [CrossRef] [PubMed]
  161. Kurata, S.; Hamada, N.; Kanazawa, A.; Endo, T. Study on Antibacterial Dental Resin Using Tri-n-Butyl(4-Vinylbenzyl)Phosphonium Chloride. Dent. Mater. J. 2011, 30, 960–966. [Google Scholar] [CrossRef] [PubMed]
  162. Kanazawa, A.; Ikeda, T.; Endo, T. Polymeric Phosphonium Salts as a Novel Class of Cationic Biocides. V. Synthesis and Antibacterial Activity of Polyesters Releasing Phosphonium Biocides. J. Polym. Sci. Part A Polym. Chem. 1993, 31, 3003–3011. [Google Scholar] [CrossRef]
  163. Chang, H.-I.; Yang, M.-S.; Liang, M. The Synthesis, Characterization and Antibacterial Activity of Quaternized Poly(2,6-Dimethyl-1,4-Phenylene Oxide)s Modified with Ammonium and Phosphonium Salts. React. Funct. Polym. 2010, 70, 944–950. [Google Scholar] [CrossRef]
  164. Kumar, V.; Malhotra, S.V. Study on the potential anti-cancer activity of phosphonium and ammonium-based ionic liquids. Bioorganic Med. Chem. Lett. 2009, 19, 4643–4646. [Google Scholar] [CrossRef]
  165. Hatakeyama, E.S.; Ju, H.; Gabriel, C.J.; Lohr, J.L.; Bara, J.E.; Noble, R.D.; Freeman, B.D.; Gin, D.L. New Protein-Resistant Coatings for Water Filtration Membranes Based on Quaternary Ammonium and Phosphonium Polymers. J. Membr. Sci. 2009, 330, 104–116. [Google Scholar] [CrossRef]
  166. Wang, K.; Zeng, Y.; He, L.; Yao, J.; Suresh, A.K.; Bellare, J.; Sridhar, T.; Wang, H. Evaluation of Quaternary Phosphonium-Based Polymer Membranes for Desalination Application. Desalination 2012, 292, 119–123. [Google Scholar] [CrossRef]
  167. Brunel, F.; Lautard, C.; Giorgio, C.D.; Garzino, F.; Raimundo, J.-M.; Bolla, J.M.; Camplo, M. Antibacterial Activities of Mono-, Di- and Tri-Substituted Triphenylamine-Based Phosphonium Ionic Liquids. Bioorganic Med. Chem. Lett. 2018, 28, 926–929. [Google Scholar] [CrossRef] [PubMed]
  168. Taladriz, A.; Healy, A.R.; Pérez, E.J.F.; García, V.H.; Martínez, C.H.R.; Alkhaldi, A.A.M.; Eze, A.A.; Kaiser, M.; Koning, H.P.; de Chana, A.; et al. Synthesis and Structure-Activity Analysis of New Phosphonium Salts with Potent Activity against African Trypanosomes. J. Med. Chem. 2012, 55, 2606–2622. [Google Scholar] [CrossRef] [PubMed]
  169. Luque-Ortega, J.R.; Reuther, P.; Rivas, L.; Dardonville, C. New Benzophenone-Derived Bisphosphonium Salts as Leishmanicidal Leads Targeting Mitochondria through Inhibition of Respiratory Complex II. J. Med. Chem. 2010, 53, 1788–1798. [Google Scholar] [CrossRef] [PubMed]
  170. Kim, T.-S.; Park, H.-D. Tributyl Tetradecyl Phosphonium Chloride for Biofouling Control in Reverse Osmosis Processes. Desalination 2015, 372, 39–46. [Google Scholar] [CrossRef]
  171. Martín-Rodríguez, A.J.; Babarro, J.M.F.; Lahoz, F.; Sansón, M.; Martín, V.S.; Norte, M.; Fernández, J.J. From Broad-Spectrum Biocides to Quorum Sensing Disruptors and Mussel Repellents: Antifouling Profile of Alkyl Triphenylphosphonium Salts. PLoS ONE 2015, 10, e0123652. [Google Scholar] [CrossRef]
  172. Worthington, R.J.; Richards, J.J.; Melander, C. Small Molecule Control of Bacterial Biofilms. Org. Biomol. Chem. 2012, 10, 7457–7474. [Google Scholar] [CrossRef]
  173. Shi, L.W.; Zhuang, Q.Q.; Wang, T.Q.; Jiang, X.D.; Liu, Y.; Deng, J.W.; Sun, H.H.; Li, Y.; Li, H.H.; Liu, T.B.; et al. Synthetic Antibacterial Quaternary Phosphorus Salts Promote Methicillin-Resistant Staphylococcus aureus-Infected Wound Healing. Int. J. Nanomed. 2023, 18, 1145–1158. [Google Scholar] [CrossRef]
  174. Guo, A.; Wang, F.; Lin, W.; Xu, X.; Tang, T.; Shen, Y.; Guo, S. Evaluation of Antibacterial Activity of N-Phosphonium Chitosan as a Novel Polymeric Antibacterial Agent. Int. J. Biol. Macromol. 2014, 67, 163–171. [Google Scholar] [CrossRef]
  175. Pinnock, T.; Voordouw, J.; Voordouw, G. Use of Carbon Steel Ball Bearings to Determine the Effect of Biocides and Corrosion Inhibitors on Microbiologically Influenced Corrosion under Flow Conditions. Appl. Microbiol. Biotechnol. 2018, 102, 5741–5751. [Google Scholar] [CrossRef]
  176. Okoro, C.C. The Biocidal Efficacy of Tetrakis-hydroxymethyl Phosphonium Sulfate (THPS) Based Biocides on Oil Pipeline PigRuns Liquid Biofilms. Pet. Sci. Technol. 2015, 33, 1366–1372. [Google Scholar] [CrossRef]
  177. Svara, J.; Weferling, N.; Hofmann, T. Phosphorus Compounds, Organic. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley: Hoboken, NJ, USA, 2000. [Google Scholar] [CrossRef]
  178. Pugachev, M.V.; Shtyrlin, N.; Sysoeva, L.P.; Nikitina, E.V.; Abdullin, T.I.; Iksanova, A.G.; Ilaeva, A.A.; Musin, R.Z.; Berdnikov, E.A.; Shtyrlin, Y.G. Synthesis and Antibacterial Activity of Novel Phosphonium Salts on the Basis of Pyridoxine. Bioorganic Med. Chem. 2013, 21, 4388–4395. [Google Scholar] [CrossRef] [PubMed]
  179. Pugachev, M.V.; Shtyrlin, N.V.; Sapozhnikov, S.V.; Sysoeva, L.P.; Iksanova, A.G.; Nikitina, E.V.; Musin, R.Z.; Lodochnikova, O.A.; Berdnikov, E.A.; Shtyrlin, Y.G. Bis-Phosphonium Salts of Pyridoxine: The Relationship between Structure and Antibacterial Activity. Bioorganic Med. Chem. 2013, 21, 7330–7342. [Google Scholar] [CrossRef] [PubMed]
  180. Dharaskar, S.; Sillanpaa, M. Synthesis, Characterization, and Application of Trihexyl(Tetradecyl)Phosphonium Chloride as Promising Solvent for Extractive Desulfurization of Liquid Fuel. Chem. Eng. Res. Des. 2018, 133, 388–397. [Google Scholar] [CrossRef]
  181. Keglevich, G.; Grün, A.; Hermecz, I.; Odinets, I.L. Quaternary Phosphonium Salt and 1,3-Dialkylimidazolium Hexafluorophosphate Ionic Liquids as Green Chemical Tools in Organic Syntheses. Curr. Org. Chem. 2011, 15, 3824–3848. [Google Scholar] [CrossRef]
  182. Sommers, K.J.; Michaud, M.E.; Hogue, C.E.; Scharnow, A.M.; Amoo, L.E.; Petersen, A.A.; Carden, R.G.; Minbiole, K.P.C.; Wuest, W.M. Quaternary Phosphonium Compounds: An Examination of Non-Nitrogenous Cationic Amphiphiles That Evade Disinfectant Resistance. ACS Infect. Dis. 2022, 8, 387–397. [Google Scholar] [CrossRef]
  183. Wang, L.; Xu, X.; Guo, S.; Peng, Z.; Tang, T. Novel Water Soluble Phosphonium Chitosan Derivatives: Synthesis, Characterization and Cytotoxicity Studies. Int. J. Biol. Macromol. 2011, 48, 375–380. [Google Scholar] [CrossRef]
  184. Qian, C.; Xu, X.; Shen, Y.; Li, Y.; Guo, S. Synthesis and Preliminary Cellular Evaluation of Phosphonium Chitosan Derivatives as Novel Non-Viral Vector. Carbohydr. Polym. 2013, 97, 676–683. [Google Scholar] [CrossRef]
Figure 1. Biofilm life cycle (reworked from the image reported in the work by Monroe et al. [74], an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium.
Figure 1. Biofilm life cycle (reworked from the image reported in the work by Monroe et al. [74], an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium.
Pharmaceutics 16 00080 g001
Figure 2. Schematic representation of the most recognized mode of action of antibacterial QASs. Symbols (+) refer to the cationic QAS, while symbols (-) indicate the negatively charged constituents of bacterial membrane.
Figure 2. Schematic representation of the most recognized mode of action of antibacterial QASs. Symbols (+) refer to the cationic QAS, while symbols (-) indicate the negatively charged constituents of bacterial membrane.
Pharmaceutics 16 00080 g002
Figure 3. Structure of gemini QASs. Two possible structures of a gemini QAS schematically represented (a); a possible general chemical structure of a gemini QAS (b).
Figure 3. Structure of gemini QASs. Two possible structures of a gemini QAS schematically represented (a); a possible general chemical structure of a gemini QAS (b).
Pharmaceutics 16 00080 g003
Chart 1. Examples of QPS chemical structures.
Chart 1. Examples of QPS chemical structures.
Pharmaceutics 16 00080 ch001
Figure 4. Number of publications per year during the last fifteen years according to Scopus and PubMed, concerning the development of QASs and QPSs for use against BFs. The survey was made by crossing the information obtained by Scopus and PubMed using the keywords Quaternary AND ammonium AND compounds AND biofilm (yellow bars) and Quaternary AND phosphonium AND compounds AND biofilm (fuchsia bars).
Figure 4. Number of publications per year during the last fifteen years according to Scopus and PubMed, concerning the development of QASs and QPSs for use against BFs. The survey was made by crossing the information obtained by Scopus and PubMed using the keywords Quaternary AND ammonium AND compounds AND biofilm (yellow bars) and Quaternary AND phosphonium AND compounds AND biofilm (fuchsia bars).
Pharmaceutics 16 00080 g004
Scheme 1. Synthesis of chitosan-based QPSs reported and prepared by Wang et al. [148]. Preparation of the quaternary phosphonium carboxylic acids (1); esterification of chitosan with the quaternary phosphonium carboxylic acids (2). EDC = 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide; NHS = N-Hydroxy-succinimide.
Scheme 1. Synthesis of chitosan-based QPSs reported and prepared by Wang et al. [148]. Preparation of the quaternary phosphonium carboxylic acids (1); esterification of chitosan with the quaternary phosphonium carboxylic acids (2). EDC = 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide; NHS = N-Hydroxy-succinimide.
Pharmaceutics 16 00080 sch001
Scheme 2. Synthesis of THPS according to Svara et al. [177].
Scheme 2. Synthesis of THPS according to Svara et al. [177].
Pharmaceutics 16 00080 sch002
Scheme 3. Synthesis of bis-phosphonium salts of pyridoxine reported by Kayumov et al. (namely compounds 4, 5, and 6) [151], as prepared by Pugachev et al. [178] and subsequently optimized [179]. Bu = butyl.
Scheme 3. Synthesis of bis-phosphonium salts of pyridoxine reported by Kayumov et al. (namely compounds 4, 5, and 6) [151], as prepared by Pugachev et al. [178] and subsequently optimized [179]. Bu = butyl.
Pharmaceutics 16 00080 sch003
Scheme 4. Synthesis of pillar [5] arene-based QPS 1 and 2 and the phosphonium monomer M reported by Joseph et al. [153]. THF = tetrahydrofuran.
Scheme 4. Synthesis of pillar [5] arene-based QPS 1 and 2 and the phosphonium monomer M reported by Joseph et al. [153]. THF = tetrahydrofuran.
Pharmaceutics 16 00080 sch004
Scheme 5. Synthesis of tributyl tetradecyl phosphonium chloride (TTPC) reported by Kim et al. [170].
Scheme 5. Synthesis of tributyl tetradecyl phosphonium chloride (TTPC) reported by Kim et al. [170].
Pharmaceutics 16 00080 sch005
Scheme 6. General procedure to prepare ATPB.
Scheme 6. General procedure to prepare ATPB.
Pharmaceutics 16 00080 sch006
Scheme 7. Synthesis of P6P-10,10 [182].
Scheme 7. Synthesis of P6P-10,10 [182].
Pharmaceutics 16 00080 sch007
Scheme 8. Synthesis of alkyl-bis-(triphenyl) phosphonium bromides (1,2-DBTPP)Br2, (1,4-DBTPP)Br2, and (1,6-DBTPP)Br2 [173]. n = 2, 4, 6.
Scheme 8. Synthesis of alkyl-bis-(triphenyl) phosphonium bromides (1,2-DBTPP)Br2, (1,4-DBTPP)Br2, and (1,6-DBTPP)Br2 [173]. n = 2, 4, 6.
Pharmaceutics 16 00080 sch008
Scheme 9. Synthesis of N-phosphonium chitosans (NPCSs) with different degrees of substitution [183]. EDC HCl = N-(3-dimetilaminopropil)-N′-etilcarbodiimide hydrochloride; HOBt = hydroxybenzotriazole.
Scheme 9. Synthesis of N-phosphonium chitosans (NPCSs) with different degrees of substitution [183]. EDC HCl = N-(3-dimetilaminopropil)-N′-etilcarbodiimide hydrochloride; HOBt = hydroxybenzotriazole.
Pharmaceutics 16 00080 sch009
Table 1. Most clinically relevant BF-producing bacteria and related infections.
Table 1. Most clinically relevant BF-producing bacteria and related infections.
Device-associatedType of DeviceTypical Bacterial SpeciesInfectionRef.
Contact lenses E. coli
P. aeruginosa
S. aureus
S. epidermidis
Candida spp.
Serratia spp.
Proteus spp.
Keratitis[56,57,58]
Central venous cathetersP. aeruginosa
Enterobacteriaceae
Klebsiella spp.
Bacteremia[57,58,59,60]
Mechanical heart valvesStreptococci
S. aureus
S. epidermidis
Bacillus spp.
Enterococci
Candida spp.
PVE[56,57,58]
Peritoneal dialysis cathetersStaphylococci
P. aeruginosa
E. coli
CESI
TI
PI
[61]
Prosthetic jointsS. aureus
CN Staphylococci
Propionibacterium spp.
Enterococcus spp.
Gram-negative bacilli
Beta-hemolytic streptococci
PPI[62]
PacemakersS. aureus
CN S. aureus
E. coli
K. pneumoniae
P. aeruginosa
A. baumannii
Endocarditis[63]
Urinary cathetersE. coli
Enterococcus faecalis
S. epidermidis
P. aeruginosa
Proteus mirabilis
K. pneumoniae
UTI[57,58,64,65]
Voice prosthesesStreptococcus mitis
Streptococcus sobrinus
S. aureus
P. aeruginosa
TTI [66]
Non-device-associatedN.E.P. aerobicus
Fusobacterium nucleatum
Periodontitis [67,68,69]
Different Bacteria
Fungi
Osteomyelitis [70]
Haemophilus influenzae
S. pneumoniae
Otitis media[71]
Fusobacterium
Streptococcus
Veillonella
Actinomyces
Prevotella
Porphyromonas
Neisseria
Eubacteria
Treponema
Lactobacterium
Haemophilus
Eikenella
Capnocytophaga
Peptostreptococcus
Leptotrichia
Propionibacterium
Staphylococcus
Dental plaque
Uropathogenic E. coli (UPEC)
K. pneumoniae
E. faecalis
UTI
P. aeruginosaCystic fibrosis
Viridans group
Streptococci
Staphylococcus spp.
Infective
endocarditis
Group A StreptococciTonsillitis
Group A StreptococciNecrotizing fasciitis
Gram-negative bacilliInfection kidney stone
UTI = Urinary tract infections; PVE = Prosthetic valve endocarditis; CESI = catheter exit site infections; TI = tunnel infections; PI = peritoneum infections; PPI = peri-prosthetic infection; CN = Coagulase-negative; TTI = tracheoesophageal tract infection.
Table 2. Nitrogen-based cationic and quaternary compounds developed in the last fifteen years endowed with anti-biofilm effects.
Table 2. Nitrogen-based cationic and quaternary compounds developed in the last fifteen years endowed with anti-biofilm effects.
Mechanism of ActionType of CompoundTarget PathogensEffects on BFsContration[Ref.] Year
Electrostatic interactions (EI)
Membrane depolarization (MD)
C12 di-cationic BODIPY-based fluorescent
amphiphiles
S. epidermidis 1
S. epidermidis 2
Prevention of BF formation
Destruction of mature BF
16 µg/mL[108] 2023
EI
Membrane disruption (MDI)
DMAHDMStreptococcus mutans⇓ Colony-forming unit counts
⇓ Metabolic activity
⇓ Exopolysaccharide synthesis
⇓ Overall acid production
⇓ Tolerance to oxygen stress (OS)
5 wt% #[109] 2019
EI, MDIQASIS. mutans
Lactobacillus acidophilus
Sortase-A protein conformational change
⇓ Carbohydrate intensities
Absence of bacterial colonies
⇓ DAPI staining
⇑ Fatty acid compositions
1–2 wt% #[110] 2020
MDI
Cell lysis
ATABS. mutans⇓ Viability of BF
⇓ Viability planktonic bacteria
5 wt% #
10 wt% #
[111] 2022
EI
Cell wall disintegration
Amphiphilic
quaternized chitosan
S. mutans29% residual BF1100 µg/mL[112] 2022
Mixed antifouling + bactericidal actionsPSBMA/PQA4C-10%, PSBMA/PQA4C-30%
PSBMA/PQA8C-10%
S. aureusPrevention of bacterial attachment and BF formationN.A.[113] 2019
Contact inhibitionDMADDMS. mutans⇓ Amount of BF
⇓ Metabolic activity
⇓ Lactic acid production
⇓ EPS
⇓ Viable S. mutans in BF
2.5–10 wt% #[114] 2020
MDI
Bind to DNA gyrase
Bind to topoisomerase IV
QA fluoroquinolonesS. aureus⇓Total biomass of P. aeruginosa ATCC 15442 BF5–10 µM[115] 2023
P. aeruginosa25 µM
MDIQASI (K21)E. faecalisDestruction of BF0.5 wt% #[116] 2021
Interference with adhesion
Interference with QS
QAL, QAHS. epidermidis
P. aeruginosa
50% ⇓ in BF formation0.037–0.15 mg/mL[117] 2019
Contact-killC14-QAAM
C14-QAMAM
S. mutans⇓ 99% in BF formation
No BF adhesion
10 wt% #[118] 2019
MDI[Ch] [Gly], [Ch] [Ala]Bacillus cereus
Pseudomonas fluorescens.
56–83% removal of biomass 70–100 mg/mL[119] 2022
⇓ 1.3–2 log CFU/cm2
culturable population
Multiple
mechanisms
Dequalinium chloride (DQC)Bacterial vaginosis
Gardnerella spp.
⇓ 80% BF biomass
⇓ 80% metabolic activity
25.64 µg/mL[120] 2021
MDDDAB *S. aureusTotal removal of BF32 µg/mL[121] 2022
P. aeruginosaDisruption of BF structure
⇓ BF polysaccharides, proteins, and phospholipids
600 µg/mL
MDDMAHDMS. mutans
S. sanguinis
Suppression of cariogenic species in BF
Modulation of BF equilibrium from cariogenic state to non-cariogenic state
3% wt#[122] 2023
Formation of eDNA–PHMG-Cl complexesPHMG-ClP. aeruginosa ATCC
S. aureus ATCC
Blockage of BF development0.1% and 0.5%[123] 2022
Interference
with the ion transport
Membrane lysis
QASIS. mutans
L. acidophilus
Actinomyces naeslundii
S. sanguis
Significant decrease in BF growth2%[124] 2020
EI, MDI, Cytoplasmic leakageDMADDMN.R.⇓ BF viability
⇓ BF formation
1.1–2.2 wt% #[125] 2023
N.R.DADMAC-based coatingsMethicillin-resistant
S. aureus (MRSA)
⇓ 99% adhesion and BF grow N.R.[126] 2018
Vancomycin-resistant Enterococcus spp. (VRE)⇓ 94% adhesion and BF grow
N.R.Guanidinium
anthranilamide
compounds
S. aureus⇓ 83–92%established BF62.4–64 µM[127] 2018
MDIQADMS. mutans(1–3 days) Significant inhibition of BF formation10 wt% #[128] 2022
N.R.Pyridine-based QASsE. coliBF eradication (MBEC)16 mg/L[129] 2018
S. aureus8–16 mg/L
EI, MDIGTMAC-modified PCL:PEG:GelMAE. coliSignificant ⇓ in CFUN.R.[130] 2022
S. aureus
S. epidermidis
EI, membrane infiltration
MDI
Metal-free quaternized carbon dotsS. aureus⇓ BF viability
⇓ BF formation
1000 µg/mL[131] 2019
MDI
DNA damage
ROS generation
MUTAB-based AuNCsB.subtilis
E. faecalis
S. pneumoniae
E. coli
P. aeruginosa
C. albicans
No live bacteria in BF of E. coli200 µg/mL[132] 2020
Hydrophobic and EI
MDI
ROS production
DNA damage
Si-QAC/TEOS NPsS. aureusEradication of mature BF
Inhibition of BF formation
100 µg/mL[133] 2022
Disrupt the charge balance of
bacterial membranes
⇑ Membrane permeability
⇑ ROS by irradiation at 660 nm
CP5/TFPP-QAE. coli
MRSA
70–80% BF dispersion100 µg/mL[134] 2022
N.R.QA polyethyleneimine PEI1200-C6E. coli
P. aeruginosa
B. amyloliquefaciens
S. aureus
>80% BF dispersion1–8 mg/mL[135] 2019
MD
MDI
phenylglyoxamide-based QA iodideS. aureus70% inhibition of BF formation
44% BF disruption
16 µM
32 μM
[136] 2021
E. coli28% BF disruption64 μM
Contact KillingDMAHDMStreptococci
S. mutans
Lactobacilli
⇓ 80% metabolic activity
⇓ 95% lactic acid production
3–5 wt% #[137] 2020
Acidic release of AgNPs
MDI by QAS
DNA damage
Ag@QAS@CMS. aureus⇓ 80% BF mass6.25 μg/mL (Ag)[138] 2020
EPS penetration ability
Acidic release of CuNPs
MDI by QAS
DNA damage
Cu@QAS@CMSulfate-reducing bacteria (SRB)⇓ 75% BF mass250 μM[139] 2022
EI, MDI, cytoplasmic leakageDMADDMS. mutans
S. sanguinis
S. gordonii
Significant ⇓ of multispecies BFs
Significant ⇓ in metabolic activity of multispecies BFs
40–200 μg/mL[140] 2019
Displacement of divalent cations from membrane
Membrane destabilization
Lysis
QA carbosilane
dendrimers and
dendron
S. aureusBF inhibition16–64 μg/mL MBIC[141] 2023
BF eradication32–64 μg/mL MBEC
MRSABF inhibition16–64 μg/mL MBIC
BF eradication16–64 μg/mL MBEC
MDI
ROS generation
Cu2+/PDDAS. aureus⇓ EPS
Kill cells in BF
144 μg/mL PDDA
26.5 μM Cu2+
[142] 2023
EI, MDI, cytoplasmic leakageN-alkylated
pyridine-based QAS
S.aureus
E. coli
52–95% inhibition of BF formation50–250 μg/mL[143] 2020
MDI
Pyridoxal-dependent enzyme targeting
Pyridoxine-based QAS of terbinafine
(KFU127)
S. aureus (SA)
C. albicans (CA)
E. coli (EC)
S. epidermidis (SE)
Full destruction of cells in BF400 μg/mL (CA)
400–800 μg/mL (CA+SA)
[144] 2020
⇓ >3 LogCFU/cm2128 μg/mL (SA)
128 μg/mL (SE)
256 μg/mL (EC)
EI, MDI, PAIN-Alkyl pyridinium QASE. faecalis85–100% killed cells in BF
No BF removal
250 µM (60′)[145] 2020
65–95% killed cells in BF
15–20% BF removal
250 µM
1′ + 10 s PAI
EI, MDI, EPS permeationPyridine-based tris-QASsE. coliMBIC
MBEC
8–32 μg/mL
16–32 μg/mL
[146] 2023
K. pneumoniaeMBIC
MBEC
16–64 μg/mL
64–500 μg/mL
S. aureusMBIC
MBEC
8 μg/mL
16 μg/mL
P. aeruginosaMBIC
MBEC
64 μg/mL
250–500 μg/mL
A. baumanniiMBIC
MBEC
32 μg/mL
250 μg/mL
C. albicansMBIC
MBEC
4–16 μg/mL
8–32 μg/mL
EI, MDI, cell lysisDTAB vs. C6P. aeruginosaFull eradication of BF (C6, 4 h)0.29 mM[147] 2018
EI, MDI, cytoplasmic leakageTMACSE. coliBF inhibition rate 57.6% 156 µg/mL[148] 2023
BF removal rate 41.6%2.5 mg/mL
S. aureusBF inhibition rate 58.5% 20 µg/mL
BF removal rate 59.01%2.5 mg/mL
EI, MDIGlut/ADBACP. aeruginosa5 log killing of sessile cells1000 ppm (2 h)
100 ppm (24 h)
[149] 2010
CDA100 ppm (2 h)
10 ppm (24 h)
Glut/ADBAC30% (24 h)–35% (2 h) biomass residual100 ppm
45% (24 h)–40% (2 h) biomass residual1000 ppm
CDA30% (24 h)–30% (2 h) biomass residual100 ppm
20% (24 h)–25% (2 h) biomass residual1000 ppm
N.R.Barquat® MB-50
Ucarcide® 42
P. aeruginosaN.R.N.R.[150] 2010
EI, MDI
EPS penetration ability due to the C18 chain
Pyridoxine-based QASsS. aureus
S. epidermidis
QAS (1–3) killed BF-detached Staphylococci cells16–32 µg/mL[151] 2015
S. aureusMBCadh (QAS 3)32 µg/mL
S. epidermidisMBCadh (QAS 3)16 µg/mL
EI by well-accessible positive charges
Host–guest properties of pillar[n]arenes
Pillar[n]arene QASs and imidalonium saltsS. aureus
E. faecalis
S. epidermidis
S. mutans
Inhibition of BF formation (MBIC50)
No biocidal, no hemolytic
0.4–8.8 µM[152] 2016
Pillar [5]arene QASsS. aureus
E. faecalis
Inhibition of BF formation (MBIC50)
No biocidal, no hemolytic
2–4 µg/mL[153] 2016
BODIPY = boron dipyrromethene; 1 ATCC 12228; 2 YT-169a; DMAHDM = dimethylaminohexadecylmethacrylate; # in dental cement or resin; ⇓ = decrease, reduction; ⇑ = increase; QASI = quaternary ammonium silane; ATAB = alkyl trimethyl ammonium bromide; D-PQA = dopamine-terminated quaternary ammonium salt polymer; D-PSBMA = dopamine-terminated poly(sulfobetaine methacrylate); DMADDM = dimethylaminododecyl methacrylate; QAL = low-molecular-weight quaternized chitosan derivatives; QAH = high-molecular-weight quaternized chitosan derivatives; ROS = reactive oxygen species; CFU = colony-forming unit; N.A. = not applicable; C14-QAAM = quaternary ammonium acrylamides with a C14 chain; C14-QAMAM = quaternary ammonium methacrylamides with a C14 chain; [Ch] [Ala] = choline alaninate; [Ch] [Gly] = choline glycinate; DDAB = didecyldimethylammonium bromide; * combined with slightly acidic electrolyzed water (SAEW); DMAHDM = dimethylaminohexadecyl methacrylate; PHMG-Cl = polyhexamethylene guanidine hydrochloride; N.R. = not reported; DADMAC = diallyl dimethyl ammonium chloride; QADM = bis(2-methacryloyloxyethyl) dimethylammonium bromide; MBIC = minimum biofilm inhibitory concentration; MBEC = minimum biofilm eradication concentration; PCL = polycaprolactone; PEG = polyethylene glycol; GelMA = gelatin methacryloyl; GTMAC = glycidyl trimethyl ammonium chloride; MUTAB = (11-mercaptoundecyl)-N,N,N-trimethylammonium bromide; Si-QAC = dimethyloctadecyl [3-(trimethoxysilyl)propyl]ammonium chloride; TEOS = tetraethyl orthosilicate; CP5 = carboxylatopillar [5]arene; TFPP-QA = tetrafluorophenyl porphyrin functionalized with a quaternary ammonium group; Ag@QAS@CM = silver-containing quaternary ammonium salt nanocomposite; PDDA = poly(diallyl dimethyl ammonium chloride); PAI = Er:YAG photoacoustic irrigation; DTAB = dodecyltrimethylammonium bromide; C6 = hexamethylene-1,6-bis-(N,N-dimethyl-N-dodecylammonium bromide; TMACS = N-(4-N′, N′, N′-trimethylmethanaminium chloride) benzoyl chitosan; Glut = glutaraldehyde; ADBAC = alkyldimethylbenzyl ammonium chloride; CDA = cocodiamine; Barquat® MB-50: contains 50% alkyl (C14 50%, C12 40%, C16 10%) dimethyl benzyl ammonium chloride (DMBA); Ucarcide® 42: contains 42.5% glutaraldehyde plus 50% alkyl (C14 50%, C12 40%, C16 10%) DMBA; MBCadh = minimum biocidal concentration of cells in biofilm; MBIC50 = the lowest concentration at which at least 50% reduction in BF formation was measured compared to untreated cells; Barquat® MB-50 and Ucarcide® 42 are from Azelis Americas, LLC, Westport, CT, USA.
Table 3. Anti-biofilm QPSs developed in the last fifteen years.
Table 3. Anti-biofilm QPSs developed in the last fifteen years.
Mechanism of ActionType of CompoundTarget PathogensEffects on BFConcentration[Ref.] Year
EI, MDI, cytoplasmic leakageTMPCS
TPPCS
E. coliBF inhibition rate 33.9% (TMPCS), 56.6% (TPPCS)156 µg/mL[148] 2023
BF removal rate 46.4% (TMPCS), 48.9% (TPPCS)2.5 mg/mL
S. aureusBF inhibition rate 53.8% (TMPCS), 62.2% (TPPCS)20 µg/mL
BF removal rate 60.4% (TMPCS), 69.9% V2.5 mg/mL
Disrupts disulfide
bonds on the cell surface
THPSP. aeruginosa5 log killing of sessile cells100 ppm (2 h)
100 ppm (24 h)
[149] 2010
35% (24 h)–45% (2 h) biomass residual100 ppm
40% (24 h)–40% (2 h) biomass residual1000 ppm
N.R.Bellacide® 350
Tolcide® PS75
P. aeruginosaN.R.N.R.[150] 2010
EI, MDI
Partial diffusion in EPS
Pyridoxine-based QPSsS. aureus
S. epidermidis
QPS (6) killed detached S. epidermidis cells
68–77% cells located in BF killed
32 µg/mL
8 µg/mL
[151] 2015
EI by well-accessible positive charges on QPSs
Host–guest properties of pillar[n]arenes
Pillar [5]arene QPSS. aureus
E. faecalis
Inhibition of BF formation (MBIC50)
No biocidal, no hemolytic
2–4 µg/mL[153] 2016
N.R.TTPCS. aureus⇓ BF (OD545 < 1)20 µg/mL[170] 2015
P. aeruginosa⇓ BF (OD545 < 1)40 µg/mL
⇓ Biofilm thickness (73.9%) and volume (73.8%)40 µg/mL
QS disruptionATPBChromobacterium
violaceum
Inhibition of BF formation
(violacein inhibition)
52.9–142.2 μM *[171] 2015
Vibrio harveyi.Inhibition of BF formation
(bioluminescence inhibition)
128.6–348.5 μM *
EI, MDI
Cell lysis, cytoplasmic leakage
TATPBP (P6P-10,10)EDR A. baumanniiBF eradication (MBEC)MBEC 32–63 µM[45] 2022
EI, MDI, MD
ROS production
(1,2-DBTPP)Br2
(1,4-DBTPP)Br2
(1,6-DBTPP)Br2
S. aureus100% inhibition of BF formation64–128 µg/mL **[173] 2023
80% ⇓ metabolic activity32–64 µg/mL **
MRSA100% inhibition of BF formation128–256 µg/mL **
80% ⇓ metabolic activity64–128 µg/mL **
EI with cell surface
Cell wall penetration
NCPSS. aureusSignificant ⇓ of BF formation64–128 µg/mL[174] 2014
E. coli64–256 µg/mL
CTPBsS. aureus
E. coli
Not activeMIC > 1600 µg/mL
N.R.THPSDesulfovibrio
Desulfomicrobium
Desulfocurvus
⇓ Microbiologically influenced corrosion17%[175] 2018
N.R.THPSSRB
APB
Significant ability to penetrate BF
100% inhibition of SRB
>85% inhibition of APB
0.6%[176] 2015
Antioxidant effects
Hydroxyl radical scavenging
ATBPB, ATPBMDR A. baumanniiN.R.6.25–25.0 μM[158] 2022
EI = electrostatic interactions; MD = membrane depolarization; MDI = membrane disruption; N.R. = not reported; TMPCS = N-(4-N′, N′, N′-trimethylphosphonium chloride) benzoyl chitosan); TPPCS = N-(4-N′, N′, N′-triphenylphosphonium chloride) benzoyl chitosan; THPS = tetrakis (hydroxymethyl) phosphonium sulfate; Bellacide® 350: contains 50% tributyl tetradecyl phosphonium chloride (TBTP); Tolcide® PS75: contains 75% THPS; TTPC = tributyl tetradecyl phosphonium chloride; ⇓ = reduction; NPCSs = N-phosphonium chitosans with different degrees of substitution (3%, 13%, and 21%); CTPB = (5-carboxypentyl)triphenylphosphonium bromide; ATPB = alkyl triphenyl phosphonium bromide; ATBPB = alkyl tributyl phosphonium bromide; * biocidal effects at concentrations ≥500 μM; EDR = extensively drug resistant; TATPBP = tetraalkyl tetraphenyl bis-phosphonium; (1,2-DBTPP)Br2 = ethyl-bis-(triphenyl)-phosphonium bromide; (1,4-DBTPP)Br2 = butyl-bis-(triphenyl)-phosphonium bromide; (1,6-DBTPP)Br2 = hexyl-bis-(triphenyl)-phosphonium bromide; SRB = sulfate-reducing bacteria; APB = acid-producing bacteria; ** refers to compound in bold; numbers in bold indicate the number of compound as reported in the study; Bellacide® 350 and Tolcide® PS75 are from Azelis Americas, LLC, Westport, CT, USA.
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

Alfei, S. Shifting from Ammonium to Phosphonium Salts: A Promising Strategy to Develop Next-Generation Weapons against Biofilms. Pharmaceutics 2024, 16, 80. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics16010080

AMA Style

Alfei S. Shifting from Ammonium to Phosphonium Salts: A Promising Strategy to Develop Next-Generation Weapons against Biofilms. Pharmaceutics. 2024; 16(1):80. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics16010080

Chicago/Turabian Style

Alfei, Silvana. 2024. "Shifting from Ammonium to Phosphonium Salts: A Promising Strategy to Develop Next-Generation Weapons against Biofilms" Pharmaceutics 16, no. 1: 80. https://0-doi-org.brum.beds.ac.uk/10.3390/pharmaceutics16010080

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