Next Article in Journal
Antiviral Agents against Flavivirus Protease: Prospect and Future Direction
Previous Article in Journal
Management of Pyrenophora teres f. teres, the Causal Agent of Net Form Net Blotch of Barley, in A Two-Year Field Experiment in Central Italy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Strategies to Combat Biofilms Using Antimicrobial Agents and Therapeutic Approaches

1
Institute of Synthetic Biology, Shenzhen Institute of Advanced Technology, Chinese Academy of Sciences, Shenzhen 518055, China
2
Department of Spinal Surgery, Shenzhen University General Hospital, Shenzhen 518055, China
3
School of Biomedical Sciences, Li Ka Shing Faculty of Medicine, University of Hongkong, Hong Kong, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pathogens 2022, 11(3), 292; https://doi.org/10.3390/pathogens11030292
Submission received: 30 December 2021 / Revised: 11 February 2022 / Accepted: 14 February 2022 / Published: 25 February 2022

Abstract

:
Biofilms are intricate bacterial assemblages that attach to diverse surfaces using an extracellular polymeric substance that protects them from the host immune system and conventional antibiotics. Biofilms cause chronic infections that result in millions of deaths around the world every year. Since the antibiotic tolerance mechanism in biofilm is different than that of the planktonic cells due to its multicellular structure, the currently available antibiotics are inadequate to treat biofilm-associated infections which have led to an immense need to find newer treatment options. Over the years, various novel antibiofilm compounds able to fight biofilms have been discovered. In this review, we have focused on the recent and intensively researched therapeutic techniques and antibiofilm agents used for biofilm treatment and grouped them according to their type and mode of action. We also discuss some therapeutic approaches that have the potential for future advancement.

1. Biofilms and Chronic Infections

Biofilms are multicellular clusters of microbes that adhere to various surfaces using an extracellular polymeric substance (EPS) made up of proteins, polysaccharides, or extracellular DNA (eDNA) [1]. The EPS provides mechanical strength, shelter from antimicrobials and host immune cells, attachment and clumping of biofilm cells, tolerance to dehydration, and assimilation of different compounds, and also provides a carbon source at nutrient-deprived conditions [2]. The capacity to form biofilms is a general feature of bacteria [3]. All eukaryotes are colonized by microorganisms that form biofilms. The biofilm microorganisms elicit definite mechanisms for early adherence to a surface, growth, and expansion of a community structure and microenvironment, and dispersal. The molecular mechanism controlling biofilm formation differs greatly amid different species and even fluctuates between separate strains of the identical species [3]. However, there are some common mechanisms for biofilm formation (see Figure 1). The microbial biofilm cycle consists of four phases: A. Initial attachment, where the planktonic cells’ initial attachment to the medical device surfaces or the host is through bacteria−surface interactions that are ultimately determined by the interplay between physicochemical interactions. B. Adherence is a stage where microbes adhere to a medical device or the host through adhesins. In this phase, microbes start to divide and form an exopolymeric substance (EPS), which improves adhesion while the formation of an EPS envelops the cells. C. Proliferation and maturation phase are whereby 3D biofilm assemblies develop where the EPS offers multifunctional and concealing microenvironments where diverse microorganisms can coexist and communicate through a quorum sensing (QS) system. D. Dispersal is the final stage where the cells escape from the biofilm to re-enter the planktonic phase [4].
Biofilm infections include both device-related and non-device-related biofilms that affect numerous people in the world each year that result in numerous deaths [5]. The microorganisms that are most frequently associated with medical devices are the staphylococci (particularly Staphylococcus epidermidis and Staphylococcus aureus), followed by Pseudomonas aeruginosa [6]. The microbes can also gain access to the host body through the contamination of therapeutic devices such as catheters, contact lenses, prosthetic heart valves and joints, voice prostheses, and pacemakers [7,8]. Biofilm production on indwelling therapeutic devices significantly affects surgical and instrumental procedures and public health as well. Wound infections, cystic fibrosis, otitis media, native valve endocarditis, bladder infections, and periodontitis are examples of non-device-related infections. The mucous layer in the host segregates bacteria from direct contact with the epithelium. Nonetheless, any rupture in the mucous layer exposes bacteria to the host epithelium and infection of mucosal surfaces [6]. To survive inside the host, the invading microbes need to overcome the epithelial wall, host-microbiome, a variety of leukocytes, and complement [2]. The immune system identifies diverse bacterial molecular patterns, though these elements could be concealed in the biofilms [2]. Biofilms decrease the efficiency of both macrophages and polymorphonuclear neutrophils [2]. This results in chronic mucosal diseases such as inflammatory bowel diseases, pharyngo-tonsillitis, rhinosinusitis, urethritis, and vaginitis [6]. The current therapeutic approaches used by medical staff include aggressive physical removal of biofilms and localized delivery of high and sustained antimicrobial chemotherapy such as antibiotics. Intravenous catheters are usually treated using a “lock therapy” which involves the treatment of a high dose of antibiotics into the lumen of the catheter for several hours [4]. Biofilms-related problems are increasing in the health care, food industry, and other fields while new antibiotics have not been produced by the pharmaceutical industry in more than a decade. Furthermore, most biofilm bacteria are tolerant to antibiotics [9]. Thus, there is an urgent need to find an alternative to antibiotics for treating biofilm-related infections.

2. Biofilm Antibiotic Tolerance

Antibiotic resistance is the acquired resistance to antibiotics through inheritable genetic mutations. In contrast, antibiotic tolerance is a physiological state of biofilm cell populations that is temporary and non-heritable [10]. Bacterial biofilms are tough to cure because of their antibiotic tolerance and might pave the way to chronic infections [11,12]. Antibiotic tolerance mechanisms in biofilm might be different from the planktonic cells [12]. The biofilm entities have a few extra tolerance mechanisms than the planktonic counterparts which hinder the treatment options and lead to the development as well as distribution of persistent bacteria [12] (Figure 2). The common causes of antibiotic failure in biofilms include: 1. Inhibition of antibiotic penetration caused by EPS barrier; 2. Accumulation of antibiotic degrading enzymes in the EPS; 3. Presence of eDNA (Tolerance to antibiofilm agents caused due to changes in the outer membrane induced by eDNA via chelating cations and resistance to antibiotics by eDNA mediated horizontal gene transfer of antibiotic resistance genes between biofilm microorganisms); 4. Quorum sensing within the biofilms that control biofilm volume, thickness, roughness, and channel formation; 5. Upregulation of efflux pumps and inactivation of drug by a. alteration of either drug or target and b. neutralization of drug; 6. Multispecies interaction and microbial diversity; and 7. Altered growth rate, stress response, and persister cells [10]. Deep inside the biofilm, dormant microorganisms known as persisters are present which are linked to a common stress response. The reduced metabolism protects the bacteria from the impacts of pH variations, osmolality, and chemical concentration [13]. Persisters are of immense clinical concern, as they play a crucial part in the antibiotic failure, relapse of bacterial infections, and serve as a pool from which resistant strains could emerge [14]. Additionally, biofilms include many concentration gradients, creating aerobic and anaerobic microenvironments which increase tolerance to antiseptics and antibiotics treatment [13]. Therefore, biofilm antibiotic tolerance is due to its multicellular nature and any interruption in step in multicellular structure might increase the antibiotic effectivity and host defenses [9].

3. Basic Strategies to Treat Biofilms

In general, two different strategies to treat biofilms exist: (1) biofilm inhibition, to prevent biofilm from forming, and (2) biofilm dispersal, to eliminate already formed biofilms (Figure 1). In order to avoid biofilm development, the adherence of planktonic cells to various surfaces or the development into premature microcolonies and also to matured biofilms should be prevented. The early attachment could be tackled in two ways: Firstly, one can modify the surface that imprints distinct 3D patterns or target physicochemical properties such as surface hydrophobicity to block microbial attachment [15,16,17]. Secondly, one can pre-condition the surfaces with chemicals to prevent initial attachment. For example, treating the surface with certain surfactants can inhibit bacterial adherence [18]. The approaches that are aimed to impede bacterial establishment on various surfaces or to avoid maturation of biofilms, typically exploit some stimuli to control the genes involved in the formation of biofilm [19] such as repressing the expressions of bacterial adhesins [20], inhibition of the biofilm EPS synthesis [21], and provoking Quorum sensing (QS) signals [22,23,24,25], or even eliminate bacteria in biofilms [26].
Preformed matured biofilms could be dispersed by breaking off biofilm assemblies and promoting detachments (Figure 1). For inducing biofilm detachment, EPS polymers should be disrupted, or cells should be programmed to disperse [17,27]. All of the antibiofilm agents and the therapeutic approaches we discuss to tackle biofilms are based on these two strategies (Figure 3).

4. Antimicrobial Agents

Antibiofilm agents belong to diverse compounds that can inhibit and eradicate biofilm formation. The established anti-biofilm compounds are chiefly extracted from natural sources while various chelating agents and synthetic compounds have been found to exhibit some anti-biofilm activity. A current review by Plakunov et al. recommended the agents to be classified into four categories based on their actions in various stages of biofilm formation (Table 1) [28]. This rule of categorization seems suitable to subcategorize a certain class of compounds but rather complicated to classify a wide range of antimicrobial compounds en bloc. In this review, we have categorized each agent based on their mode of action and their type. Here, we review some contemporary antibiofilm agents and the therapeutic approaches that can be used as alternatives to antibiotics for treating biofilms.

4.1. Surface Attachment Inhibitors

The initial phases of attachment are very crucial in the biofilm development process. The control of surface attachment can inhibit the whole process of biofilm development. Biofilm formation can also be inhibited by the inhibition of adhesin and EPS molecules. When the bacteria have short-range interaction with the surface, the hydrophobic interactions, dipole, ionic, and hydrogen bonds begin to dominate over some other interactions, and then bacteria start to attach through the cellular or molecular phase [29] (Figure 1). Surfactants are the popular choice of antimicrobial agents for inhibiting bacterial adhesion to the surfaces as they decrease the interfacial tension between two substances. Surfactants are amphiphilic as they comprise of hydrophilic and hydrophobic moiety and at the same time, they can be categorized as non-ionic, anionic, cationic, and amphoteric surfactants [27]. Triton X-100 and Tween 80 (Polysorbate 80) are the two popular non-ionic, synthetically derived, and regularly used surfactants in laboratories (Table 2). Tween 80 decreased S. aureus medical device associated biofilm development at concentrations safe in humans [30]. Triton X-100 could stimulate autolysis by enhancing bacterial vulnerability to antibiotics and altering the architecture and physiological features of biofilms by reducing the protein and carbohydrate constitution in the EPS [31,32]. Biosurfactants are the surface-active compounds created by microorganisms that comprise structurally diverse biomolecules [33]. Cationic surfactants such as quaternary ammonium compounds (QACs) are used as disinfectants within the food industry and used in several medical conditions. QACs bind to negatively charged areas in microbes that cause stress to the cell wall, lysis, and cell death. QACs could also cause protein denaturation that affects cell wall permeability and reduces the uptake of nutrients. Non-ionic-based surfactants incorporating poloxamer 188, are regarded as non-cytotoxic and therefore represent a useful combination in wound care. Numerous studies using antimicrobials together with poloxamers showed enhanced antimicrobial efficacy [34]. Similar to regular surfactants, numerous biosurfactants have antimicrobial activities while some even seem to prevent surface colonization by pathogens [27]. One of them is rhamnolipid, which is the chief glycolipid formed by several bacterial species, chiefly by P. aeruginosa [35], and stimulates biofilm dispersal in P. aeruginosa, S. aureus, Salmonella enteritidis, and Listeria monocytogenes [36]. Furthermore, rhamnolipids from P. aeruginosa W10 were also known to disperse biofilms of various industrial bacterial strains on the pipelines [37]. Since biosurfactants are usually associated together with isomers and cogeners and rarely in pure form, the purification process could be exhaustive and expensive [38]. At the same time, they could be very cytotoxic and hemolytic due to their activity with cellular membrane [39]. Therefore, their use for controlling biofilms could be limited to coating medical devices and anti-adhesive agents.

4.2. Compound Inducing Cell Lysis

To inhibit the biofilm formation process, it is best to kill the bacteria in the earlier phases of biofilm development. It can be achieved through targeting the cellular components and mechanisms. The breakdown of peptidoglycan that makes the cell wall of bacteria could inhibit biofilm formation as it changes the constitution of teichoic acids and proteins on the cell wall and likewise releases the signals that regulate genes related to biofilm [54]. Enzymes such as transglycosylase and peptidoglycan hydrolases (endolysins) break the cell wall and often result in bacterial cell death [55,56].
Cell division is critical for bacterial existence in the biofilms and for spreading further to new surfaces [54]. Chelating agents such as Ethylenediaminetetraacetic acid (EDTA) can damage the cell wall, subsequently disrupting the biofilms via sequestering zinc, magnesium, iron, and calcium [40]. EDTA is generally safe for use in prescription medicine and in small amounts in food preservatives. Similarly, Chitosan is a natural polymer used in numerous applications in the biomedical field because of its biodegradability, bioadhesive property, and bioactivity [56]. It is known to disrupt negatively charged cell membranes due to its cationic nature (Table 2) [41]. Therefore, by using such agents we can tackle the bacteria in the early phase of biofilm development.

4.3. Antiquorum Sensing Molecules

QS is a cell–cell interaction mechanism in the microbial groups to harvest a coordinated effort on regulating the genes related to virulence, biofilm formation, antibiotic tolerance, and survival. The QS mechanism is brought by a set of two proteins in Gram-negative bacteria. One of these is an autoinducer (AI) which leads to signaling molecule production, while the other one acts in response to the AI [57]. In contrast to Gram-positive bacteria that use secreted autoinducing peptides (AIPs) such as AIs for QS, Gram-negative bacteria employ homoserine lactones (HSL). In addition, unlike Gram-negative bacteria that use a regulator-type protein as an autoinducer sensor, Gram-positive bacteria employ two-component adaptive response proteins for sensing AIs. This mechanism of signaling is stimulated by a phosphorylation/dephosphorylation cascade [58]. Cyclic di-GMP (c-di-GMP) is also known as a secondary messenger since it is responsible for controlling several functions such as motility, cell cycle, differentiation, developmental transitions, adhesion, aggregation, biofilm production, and virulence in various pathogens. In several Gram-negative bacteria such as Escherichia coli, Gluconacetobacter xylinus, P. aeruginosa, and Salmonella enterica, c-di-GMP plays a significant role in the shift in between motile to sessile lifestyle, construction of three-dimensional (3D) biofilms, and during biofilm dispersal [59,60,61].
Many natural and synthetic compounds act as anti-QS molecules that target various QS signaling molecules (Table 3). A natural compound such as garlic was able to reduce the virulence factors progression and decrease the QS signal production in P. aeruginosa in a mouse urinary tract infection (UTI) model [62]. In another study by Persson et al., it was reported that garlic extracts inhibit biofilm production in six clinical isolates of bacteria [57]. Besides, through the rigorous design and screening, all the biological compounds from a potent QS inhibitor interrupted QS signaling by negatively regulating the transcriptional regulators LasR and LuxR [63]. Ichangin and isolimonic acid are the strong regulators of cell–cell signaling in bacteria, while they are the effective repressors of biofilm and the type III secretion system. Moreover, isolimonic acid also seems to affect AI-3/epinephrine generated cell–cell signaling pathways in QseA and QseBC dependent manner [64,65]. It also interfered in AI-2 based QS by reducing the LuxR DNA-binding potential in several Vibrio spp. [65]. Another natural compound cinnamaldehyde was known to reduce E. coli swimming motility and alter biofilm structure and formation [66]. It was also found that hordenine, a potent phenylethylamine alkaloid obtained from barley, exhibits a dose-dependent decline in the production of the signaling molecule and affects biofilm production in P. aeruginosa [67]. Furthermore, hordenine also effectively reduces QS-associated gene expression and virulence factors of P. aeruginosa PAO1 [67,68]. This suggested that hordenine appears to be a novel anti-QS agent that could protect from pathogens [67]. Plant polyphenols known as quercetin are reported to significantly reduce biofilm production and other virulence factors at a lower concentration than formerly known substances and plant extracts [69,70,71,72]. Furthermore, a study on QS-associated transcriptional changes revealed that LasI/R, RhlI/R expression levels involved in QS were significantly decreased [73]. Autoinducing peptide type I (AIP-I) stimulated MRSA biofilms dispersal on titanium disks, causing detached MRSA more vulnerable to treatment with rifampin and levofloxacin [74]. RNAIII-inhibiting peptide (RIP) resulted in a 7-log reduction in MRSA in a mouse wound model [75]. While the increased effectiveness of antibiotic treatment with QSI in vivo is promising, reduced bacterial loads often depend on the strain and biofilm model [4].
Naturally, nitric oxide (NO) is recognized as the universal signaling molecule that can circulate easily in biological systems. However, some studies highlight the role of NO in biofilm dispersal by targeting the QS system of bacteria [27]. The NO producing agents including sodium nitroprusside (SNP) induced lifestyle transition in bacteria, from the sessile biofilm state to a mobile planktonic state by reducing the amount of intracellular c-di-GMP, thereby causing dispersal of P. aeruginosa biofilms [76]. A similar effect of biofilm-dispersal by NO donors has been verified in Bacillus subtilis as well [77]. These studies underscore that NO generating agents could be potential antibiofilm agents.

4.4. Synthetic Small Organic Molecules

The design of synthetic small organic molecules has paved a new route to overcome antibiotic tolerance and interfere with biofilms [78]. It has drawn remarkable attention in the past few decades. Numerous research shows that small organic molecules inhibit biofilms by different modes of action (Table 4).
Some imidazole and benzimidazole compounds are able not only to inhibit biofilms but also to disperse them. The molecular mechanism behind the antibiofilm effect for the 5-phenyl-2-aminoimidazole was interpreted in Salmonella typhimurium [79]. The study emphasizes the potential of 5-phenyl-2-aminoimidazole to decrease the expression of CsgD, and adrA and csgB genes regulated by it, thus preventing the biofilm EPS formation [79]. Sambanthamoorthy et al. synthesized a 5-methoxy-2-[(4-methyl-benzyl) sulfanyl]-1H-benzimidazole, commonly known as ABC-1 (antibiofilm compound-1) which showed an antibiofilm effect against the Gram-negative bacteria Vibrio cholera and P. aeruginosa [80]. ABC-1 was also able to inhibit biofilms in Gram-positive pathogens including S. aureus at lower concentrations by targeting eDNA, polysaccharide intercellular adhesion (PIA), and Protein A (SpA) expression [1]. Likewise, Frei et al. confirmed the strong effect of the 5, 6- dimethoxy-2-aminobenzimidazole not only inhibited P. aeruginosa biofilms but also dispersed them by targeting two QS receptors, LasR and RhlR [81].
Pyrazole is an exceptional aromatic heterocyclic compound with five-membered rings, is also known to be a biofilm modulator. Suresh et al. tested three Pyrazolo-pyrimido [4,5-d] pyrimidines (compounds 19 a–c) and found that it was able to inhibit Gram-positive bacteria including S. aureus, B. subtitlis, and Microococcus luteus [82]. Remarkably, biofilm treatment with compound 19b displayed a substantial surge in intracellular ROS levels in M. luteus at the dose of 0.5 μg/mL, which caused the cells to undergo oxidative stress that caused membrane damage leading to cell lysis and death [83,84].
Indole derivatives are known to repress motility, chemotaxis, and adhesion in E. coli. In a screening of six plant and animal derivatives of indole, indole-3-carboxaldehyde and 3-indolylacetonitrile were found to be potential biofilm inhibitors against P. aeruginosa and E. coli O157: H7. These compounds decrease biofilms by reducing curli production without affecting microbial growth [85,86].
2-Phenylhydrazineylidene derivatives can prevent bacterial adhesion by Sortase A (SrtA) inhibition [87], a transpeptidase enzyme that aids in biofilm production by incorporating cell-surface proteins into the Gram-positive bacteria cell wall. Inhibition of SrtA is also associated with loss of virulence factors in S. aureus, including attenuation in the binding potential to fibrinogen, and fibronectin, lgG, along with a decrease in biofilm formation [1,88,89]. Pyrrole derivatives are also found to inhibit biofilms in Gram-positive pathogens. For example, Dihydro-pyrrol2-ones (DPO) derivatives such as diethyl1-(3-chlorophenyl)-4-((3-chlorophenyl) amino)-5-oxo-2,5-dihydro-1H-pyrrole-2,3-dicarboxylate exhibited inhibition in P. aeruginosa growth and biofilm formation by inhibiting mannitol dehydrogenase (MDH) and eDNA. MDH is involved in the synthesis of alginate which is one of the EPS components of P. aeruginosa [90].
Brominated furanone derivatives are known to inhibit biofilms in different bacterial species. The synthetic (Z)-5-bromomethylene-2(5H)-furanone repressed microbial communication mediated by AI-2 in several Streptococci such as Streptococcus angionus, Streptococci intermedius, and Streptococcus mutans. Similarly, bicyclic brominated furanones inhibited AI-2 mediated QS in Tannerella forsythia, Porphyoromonas gingivalis, and Fusobacterium nucleatum [91].
Halogenated phenazines showed powerful activity against Methicillin-resistant S. epidermidis (MRSE), Methicillin-resistant S. aureus (MRSA), and Vancomycin-resistant Enterococci (VRE) by binding with iron (II) and copper (II) that exhibited antibiofilm activity [92]. In the past, several bacterial infections were treated with quinolones. However, using an in silico virtual screening method, it has been lately identified that quinolone compound Ia could reduce P. aeruginosa biofilms by inhibiting PqsR (regulatory proteins). It also showed a synergistic effect with other antibiotics such as tobramycin [93]. Sommer et al. showed the 3,4-dimethoxycinnamide derivative showed biofilm inhibition in P. aeruginosa by inhibiting LecB [94].
Although several small molecules have proved to be efficient biofilm inhibitors, none of these agents have reached clinical use due to a lack of experiments in animal models. Therefore, new in vivo studies using small molecules are urgently needed to assess their therapeutic potential [78].
Table 4. Small organic molecules with the known biofilm inhibition or eradication mechanism.
Table 4. Small organic molecules with the known biofilm inhibition or eradication mechanism.
Small Organic MoleculesBiological RoleEffective AgainstType of CompoundReference
15-phenyl-2-aminoimidazoleReduction in transription of CsgD, csgB and adrAS. typhimuriumImidazole derivative[79]
2ABC-1Reduction in SpA and PIA production and decrease eDNA releaseS. aureus, Gram-positive and Gram-negative pathogensImidazole derivative[1,80]
35,6-dimethoxy-2-aminobenzimidazoleReduction in QS receptors (LasR and RhlR)P. aeruginosaImidazole derivative[81]
4Pyrazolo-pyrimido [4,5-d] pyrimidinesROS accumulation, loss of membrane integrityM. luteus, S. aureus, B. subtilis, E. coli, K. planticolapyrazole compound[82,83,84]
5Indole-3-carboxaldehyde and 3-indolylacetonitrileReduction of curli productionE. coli, P. aeruginosaindole and carbazole derivative[85,86]
6Phenylhydrazine analoguesSrtA inhibitionS. aureus2-Phenylhydrazineylidene derivatives[87]
7Diethyl 1-(2-chlorophenyl)-4-((3-chlorophenyl)amino)-5-oxo-2,5-dihydr-1H-pyrrole-2,3-dicarboxylateMDH and eDNA inhibitionP. aeruginosaPyrrole derivatives[90]
8(Z)-5-bromomethylene-2(5H)-furanoneAI-2-mediated QS inhibitionS. anginosus, S.intermedius and S. mutansFuranone and oxazolidinone derivatives[91]
9Bicyclic brominated furanonesAI-2-mediated QS inhibitionF. nucleatum, P. gingivalis and T. forsythiaFuranone and oxazolidinone derivatives[91]
10Halogenated phenazineBind with copper (II ) and iron (II)MRSA, MRSE, and VREPhenazine and quinolone derivative[92]
11Quinolone compound IaPqsR inhibitionP. aeruginosaQuinolone derivative[93]
123,4-dimethoxycinnamide derivativeLecB inhibitionP. aeruginosaCinnamide derivative of d-mannose[94]

4.5. Secondary Metabolites

Secondary metabolites (SM) do not directly contribute to the basal metabolism of its producing organism instead act as essential factors to either attract, repel, or kill other organisms and thereby increase the chance of self-survival [42,95]. Unique secondary plant metabolites such as Citrus limonoids presented their potential to affect biofilm formation and cell–cell signaling in Vibrio harveyi by modulating the expression LuxO, but not the promoter activity of LuxR (Table 2) [42].
Since marine organisms are a rich source of novel bioactive metabolites, studies on marine fungal and bacterial secondary metabolites have been gradually growing for the development of novel therapeutic agents [96]. For example, a secondary metabolite identified as cyclo(l-Tyr-l-Leu) produced from a marine ascomycete Penicillium spp isolated from the sponge Axinella corrugata which inhibited biofilm formation by S. epidermidis [43]. Similarly, Park et al. described the discovery of three novel secondary peptidic metabolites known as cahuitamycins from Streptomyces gandocaensis that were the inhibitors of Acinetobacter baumannii biofilms [44]. Thus, secondary metabolites could be abundantly found where the microorganisms coexist together. This highlights that nature could be an unlimited source for drug discovery.
Seaweed secondary metabolites such as phlorotannin possess antibacterial properties. Several in vivo studies and clinical trials exist on the health benefits of Phlorotanin. However, these studies were not based on the antibiofilm properties of phlorotannin [45,97]. The current studies in vitro show antibiofilm properties of secondary metabolites, but more in vivo studies on the antibiofilm properties seem to be required.

4.6. Antibiofilm Peptides

Antimicrobial peptides (AMPs) are a type of innate defense mechanism in different eukaryotes which was first discovered by Kiss and Michl in the 1960s. AMPs are cationic and hydrophobic residues containing molecules that can interact with various microorganisms such as bacteria, fungus, protozoa, and some enveloped viruses [98,99,100]. Since AMPs have low antigenicity and rapid killing effect in comparison to conventional antibiotics, it has been showing momentous potential in recent years [101]. Studies show sub-minimal inhibitory concentration (MIC) of some AMPs is also able to inhibit biofilm in various pathogens, thus these peptides are termed antibiofilm peptides (ABPs) [102]. Antimicrobial peptides display a wide range of antibiofilm effects by (1) cleavage of peptidoglycan, (2) change of membrane permeabilization or membrane potential, (3) neutralization or disassembly of lipopolysaccharides, (4) inhibition of cell division and cell survival, (5) modulate the synthesis of adhesion molecule synthesis and function, and (6) repression of the stringent response of the bacteria [54,103]. Some ABPs and their mechanisms for biofilm inhibition or dispersal are given in Table 5.
AMPs such as nisin and bovicin HC5 reduced the S. aureus adhesion to the polystyrene surfaces and altered the cell as well as polystyrene surface hydrophobicity. These AMPs also changed the biofilm-related gene expressions in planktonic cells [104]. On the other hand, a popular human cathelicidin AMP, LL-37, and indolicidin were known to inhibit P. aeruginosa biofilm formation. It was likely achieved by inhibiting the transcription of Las and Rhl QS systems [105,106]. LL-37 also inhibits biofilm formation of P. aeruginosa by upregulating the expression of genes needed for type IV pili biosynthesis and function [105]. Likewise, peptide 1037 efficiently inhibits biofilm production by the Gram-positive L. monocytogenes and Gram-negative pathogens Burkholderia cenocepacia and P. aeruginosa. Peptide 1037 directly reduced biofilm production by decreasing swarming and swimming motilities, inducing twitching motility, and downregulation of several genes related to biofilm production [106].
Fuente Nunez et al. found an effective anti-biofilm peptide 1018 which could bind and degrade (p)ppGpp, a crucial signal required for the formation of biofilm [102]. Peptide 1018 treatment completely prevented biofilm production at much lower concentrations, which did not alter planktonic growth and also caused degradation of mature biofilms in E. coli, A. baumannii, P. aeruginosa, Klebsiella pneumoniae, MRSA, S. Typhimurium, and Burkholderia cenocepacia [102]. D-enantiomeric protease-resistant peptides DJK-5 and DJK-6 could reduce (p)ppGpp in biofilms of P. aeruginosa to a greater amount than 1018 [107].
Apidaecin, pyrrhocoricin, and drosocin are 18–20 amino acid, proline-rich residues that were initially obtained from insects. These AMPs attack a target microbial protein in a stereospecific manner. They interact with the bacterial heat-shock protein DnaK by inhibiting chaperone-assisted protein folding and limiting the DnaK ATPase activity [108,109,110]. The peptide antibiotic microcin B17 (MccB17) is the first peptide known to repress a type II DNA topoisomerase activity. MccB17 blocks E. coli DNA gyrase by trapping an enzyme-DNA cleavable complex [111]. PR-39, an AMP which was obtained from the upper portion of a pig’s small intestine, could kill growing bacteria faster than non-growing cells. It is suggested that PR-39 kills bacteria by stopping protein and DNA synthesis [112].
Table 5. Antibiofilm peptides with known modes of actions.
Table 5. Antibiofilm peptides with known modes of actions.
BiofilmsAMPAmino Acid SequenceMW (g/moL)No. of ResiduesMode of ActionReferences
P. aeruginosaLL-37LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES4493.3337Reduces swimming and swarming motilities, promotes twitching motility, downregulates the genes required for biofilm formation and influences QS system[105,106]
1037KRFRIRVRV1229.549[106]
1018VRLIVAVRIWRR1536.9312binds and degrades (p)ppGpp[102]
Esculentin-1a (1–21)GIFSKLAGKKIKNLLISGLKG2185.7321Disrupts cell membrane[113]
RN3(5-17P22-36)RPFTRAQWFAIQHISPRTIAMRAINNYRWR3758.3830Depolarizes and permeabilize cell membrane[114]
S4 (1–16)ALWKTLLKKVLKAAAK1782.2916disintegrates and release membrane lipids[115]
DJK-5VQWRAIRVRVIR1551.9112degrade (p)ppGpp[107]
S. aureusNisin AMSTKDFNLDLVSVSKKDSGASPR3354.123Depolarizes cell membrane[116]
lacticin QMAGFLKVVQLLAKYGSKAVQMAWANKGKILDWLNAGQAIDKVVSKIKQILGIK5785.0553
Nukacin ISK-1KK-KSGVIPTVSHGCHMNSFQFVFTCC2886.4426
HC5VGXRYASXPGXSWKYVXF1616.8414alter surface hydrophobicity[104]
S. epidermidisHepcidin 20ICIFCCGCCHRSHCGMCCKT2208.820Acts on Polysaccharide Intercellular Adhesin[117]
Human β defensin 3 (HBD-3)GIINTLQKYYCRVRGGRCAVLSCLPKEEQIGKCSTRGRKCCRRKK5161.2445targets icaA, icaD, and icaR genes[118]
S. mutansP1PARKARAATAATAATAATAATAAT215824interferes and degrades EPS[119]
E. coliPyrrhocoricinVDKGSYLPRPTPPRPIYNRN2340.6720bind with DNaK[108,109,110]
ApdidaecinGNNRPVYIPQPRPPHPRI2108.4418
DrosocinGKPRPYSPRPTSHPRPIRV2198.5619
Microcin B17VGIGGGGGGGGGGSCGGQGGGCGGCSNGCSGGNGGSGGSGSHI3255.3543Inhibition of DNA replication by inhibiting type II DNA topoisomerase[111]
PR-39RRRPRPPYLPRPRPPPFFPPRLPPRIPPGFPPRFPPRFP4720.739stop the synthesis of DNA and protein[112]
Some antimicrobial peptides polarize and permeabilize the cell membrane and finally cause membrane disruptions. Esculentin (Esc (1–21) is the AMP derived from the frog skin. It could breach the cytoplasmic membrane of Gram-negative bacteria for example P. aeruginosa PAO1 inside the biofilms and lead to a discharge of β-galactosidase. However, this impact on biofilm cells was slower when compared to the planktonic cells [113]. Similarly, another synthesized peptide, RN3(5-17P22-36) was acquired from the cationic proteins of eosinophil granules. It was also capable of causing cell lysis by membrane disruption in biofilms. There was a 2–3-fold reduction in membrane depolarization in the biofilm cells as compared with the planktonic cells at the same dose of AMP [114]. Another peptide, known as S4 (1–16) M4Ka, which is a derivative of dermaseptin S4 also causes the flux of membrane lipids, bacterial dispersal, thus leading to a reduction in P. aeruginosa biofilm [115]. Bacteriocins such as nukacin ISK-1, lacticin Q, and nisin A also destroyed the membrane potential of S. aureus biofilm cells and caused the ATP flux from the cells [116].
Several AMPs also target biofilm EPS components. AMPs such as P1 could destroy biofilm EPS formed by Streptococcus mutans that resulted in decreased biofilm formation on polystyrene and saliva-coated hydroxyapatite [119]. Another AMP, hepcidin 20, was able to decrease EPS load and affect the biofilm structure of S. epidermidis by inhibiting PIA [117]. Similarly, Human β -defensin 3 (HBD-3) induced downregulation of ica operon genes, such as icaA and icaD, and at the same time upregulated icaR genes of S. epidermidis thus decreasing PIA production resulting in biofilm inhibition [118]. As all of these results suggest antibiofilm peptides can control biofilms using several different kinds of mechanisms, these peptides can be further classified into several sub-groups as suggested by Plakunov et al. [28].
A few antibiofilm peptides have even been used in treating biofilms in animals and agriculture. Since antibiofilm peptides have a complex mechanism of action, improving our understanding of these mechanisms in biologically relevant situations appears to be important for determining structure–function associations and finally optimizing synthetic antibiofilm peptides for improved antibiofilm potential [120].

4.7. Compounds Targeting Metabolism

Studies into bacterial metabolism show that certain metabolites are necessary for biofilm formation and stability [121]. Recent reports showed small-molecule metabolites and correlated metabolism were essential for biofilm development and dispersal. Pisithkul et al. found that in the early phases of biofilm development, tricarboxylic acid (TCA) cycle activity was increased, iron metabolism and the transport was reorganized, a metabolic shift had occurred from fatty acid biosynthesis to fatty acid degradation, and a switch from acetate to acetoin fermentation took place in B. subtilis [122]. In another study by Lu et al., they first evaluated the difference in metabolism between the biofilm and planktonic populations of UTI89 (uropathogenic E. coli, UPEC strain) by using mass spectrometry-based targeted and untargeted metabolomic methods, together with cytological imaging, that enabled in identifying the targeted metabolites and related metabolic pathways involved in biofilm development [123]. Interestingly, they could also find distinct changes in both metabolism and phenotypic morphology in two patterns. Moreover, they recognized and categorized 38 differential metabolites and three of the associated metabolic pathways, namely carbohydrate metabolism, amino acid metabolism, and glycerolipid metabolism, were changed typically during biofilm production [123]. In a different study, tea tree oil showed antimicrobial and antibiofilm activity against S. aureus (Table 2) and also changed its metabolism by dramatically affecting the expression of genes associated with the pyrimidine metabolism pathway, purine metabolism pathway, glycine, serine, and threonine metabolism pathway, and amino acid biosynthesis pathway [52]. Some research also showed that treatment with exogenous amino acids for example L-arginine was able to control the biofilm formation by repressing the genes that are essential in the formation of S. mutans biofilm EPS [124]. Hence, from the recent knowledge about biofilm metabolism, key biofilm metabolites and the chief metabolic pathways could be sorted out. Metabolic engineering into these pathways could be the next essential approach in tackling biofilms.

4.8. EPS Degrading Enzymes for Biofilm Dispersal

Degradation of the EPS by EPS-degrading enzymes such as α-amylase, Dispersin B (DspB), and DNase I is a popular antibiofilm strategy (Table 2). The damage to the basic biofilm component permits more infiltration of antibiotics, thus improving the efficiency of the antibiotic. α-amylase, DspB, and DNase I degrade exopolysaccharides, biofilm EPS, and eDNA, respectively [51,125], which decreases biofilm production as well as degrades mature bacterial biofilms such as Vibrio cholerae, S. aureus, and P. aeruginosa [46]. Polyamine norspermidine and D-amino acids are some of the naturally produced small molecules by bacterial communities that induce the mature biofilm dispersal and also prevent biofilm production in E. coli and S. aureus [47,48,126]. A different study reported that N-acetylcysteine/NAC and Tween 80 discretely or together with other antibiotics could effectively disperse non-pigmented rapidly growing mycobacteria (RGM) biofilms [49]. Similarly, the researchers in Japan reported that Esp, a serine protease produced by S. epidermidis can prevent as well as disperse the S. aureus biofilms in vitro. It could likewise repress S. aureus nasal colonization in vivo [50]. A similar study by Park et al. also showed that the proteases from P. aeruginosa were able to prevent biofilm production and induce dispersal in S. aureus [53]. Thus, EPS-degrading enzymes have the potential to be used in biofilm dispersal strategy as an antimicrobial agent.

5. Other Therapeutic Approaches

5.1. Phage Therapy and CRISPR-Cas 9

Bacteriophage therapy has been used for over 50 years to treat bacterial infections. Recent experimental and clinical studies have demonstrated a remarkable impact on treating both wound biofilm infections as well as device-related infections [127]. The existence of huge numbers of bacteria in the biofilms enables the bacteriophages to rapidly and efficiently infect their host and subsequent multiplication of the bacteriophage. Bacteriophages have many qualities that could make biofilms vulnerable towards them. They are known to produce or induce enzymes that degrade the ECM. They are also able to infect persister cells. Tkhilaishvili et al. have shown that bacteriophages such as Sb-1 enhanced antibiotic activity against biofilm (Table 6) [128]. Moreover, it also degraded the EPS, mainly the polysaccharide content in the biofilms, and targeted the persister cell of S. aureus [128]. It is perhaps predictable that bacteriophages can also target the bacteria in biofilms since they prey on bacteria naturally [129]. One of the merits of phage therapy over antibiotic therapy is that the treatment is much more specific. A phage attaches to one specific bacterial strain leaving others unharmed in contrast to the use of antibiotics, which may kill not only harmful bacteria but also helpful bacteria that live in our gut. At the same time, the specificity of phage therapy could also turn into a disadvantage as the matured and naturally produced biofilms could be phage resistant [130].
Clustered regularly interspaced short palindromic repeats and CRISPR-associated protein 9 (CRISPR-Cas9 system) is a novel technology that has been applied for genomic editing in some prokaryotes and eukaryotes [131]. Most recently Zuberi et al. have knockdown the luxS gene and fimbriae-associated gene (fimH) using the CRISPRi technology for regulating biofilm-associated infections [132,133]. Similarly, Hegde et al. employed the CRISPR/Cas9 system to make an ompA (outer membrane protein A) gene knockout in a dominant mosquito microbiota Cedecea neteri [134]. The ompA mutant displayed a reduction in biofilm formation capability, significantly decreased infection in adults, and was also found to be less prevalent in adults [134]. Thus, CRISPRi seems to be another promising technology for controlling biofilms.
Table 6. Some therapeutic approaches effective against biofilms.
Table 6. Some therapeutic approaches effective against biofilms.
Type of TherapyTargetBiofilmType of AssayReferences
Bacteriophage therapyEPS polysaccharideS. aureusin vitro[128]
Vaccine (Staphvax)capsular polysaccharide serotypes (CP5 and CP5)S. aureusphase III clinical trials[135]
CRISPR/CasluxS and fimH genesE. coliin vitro[132,133]
CRISPR/CasompAC. neteriin vivo[134]
Photodynamic therapyEPSS. aureusin vitro[12]
Virus like particlesagr QS systemS. aureusin vivo[135,136]
Vaccinepili-S and integration host factor (IHF)H. influenzaein vivo[137]
Catalytic antimicrobial robotskill cells and detach biofilmsS. mutansin vitro[138]
Nitric oxide-releasing nanoparticlesN/AS. aureusin vivo[139]
Arikayce™inhibition of protein synthesisM. aviumphase III clinical trials[140]
Fluidsomes™inhibition of protein synthesisP. aeruginosaPhase II clinical trials[140]
Calcium fluoride nanoparticlesvarious cellular processesS. mutansin vitro and in vivo[141]

5.2. Vaccines

Antibiotic-resistant species also evolve to colonize and evade host immunity. The approach of targeting biofilms using vaccines poses numerous difficulties in targeting bacterial biofilms since vaccines are specific to a particular microorganism and show large variability in the expression of vaccine-targeted epitopes [4]. Nevertheless, recently several conjugative vaccines have been formulated that target typically conserved EPS components. Staphvax is a conjugative vaccine that comprises polysaccharides and protein that acts on capsular polysaccharide serotypes (CP5 and CP5) in S. aureus [135]. The DNABII family of DNA-binding proteins, which consists of the integration host factor (IHF), gives structural integrity to eDNA in biofilms [142]. Studies confirm that antibodies against E. coli IHF can bind to DNABII in many bacterial species and destabilize biofilms by releasing individual bacterium. DNA II, when combined with antibiotic therapy, showed a synergistic effect in treating biofilms of numerous bacterial species such as oral bacteria [143], UPEC [144], P. aeruginosa [145], and S. aureus [146,147] in a murine lung infection model [142]. The next approach is to combine DNABII antibodies with vaccines. A study using IHF and recombinant soluble type IV pili co-administered with an adjuvant in an animal model of otitis media with nontypeable Haemophilus influenzae (NTHi) showed early NTHi eradication and prevention of disease [137]. Alternatively, to detect immunogenic mimic of a QS peptide, the usage of Virus-like Particles (VLPs) is increasing. Some results indicate that a diagnosis of VLP-based epitope for the development of a vaccine that targets agr signal disruption could be effective against S. aureus SSTI [135,136]. Better molecular identification of biofilm-associated genes can advance vaccines development against bacterial infections in the future [135].

5.3. Biomaterials and Nanoparticles

Pathogenic biofilms formed on implantable medical devices (IMDs) or human tissues have caused a huge risk in global healthcare. Functionalized biomaterials could be a novel approach to combat and eradicate pre-existing biofilms. Adherence of bacteria is the initial stage of IMD-associated infections that enables bacteria to colonize in the implants. The inhibition of biofilms could be achieved by coating the implants with biomaterials with antifouling and antibacterial properties. Biofilm eradication could be achieved by using nanoparticle (NP)-coated drugs to disperse biofilms [148].
NPs are regarded as a substitute for antibiotics for combating multidrug-resistant and biofilm-associated infections [149]. The biofilm-NP interaction is a three-step process: (1) transport of NPs around the biofilm, (2) attachment of NPs to the biofilm EPS, and (3) penetration of NPs into the EPS and migration within the biofilm through diffusion which might be dependent on the biofilm pore sizes, the charges, hydrophobicity, and the EPS chemical gradient [150]. Drawbacks of antibiotic therapy, such as reduced penetration into the biofilm, could be overcome easily through their nano-formulations that can cross the biological barrier.
Naturally forming and engineered NPs could differ largely in their physicochemical properties including size, shape, and charge [150]. For the last few years, different types of NPs have been used as antibiofilm and antimicrobial metal NPs, organic NPs, green NPs, and their combinations [151]. Several reports exist on NP-based eradication of biofilm communities [152,153,154]. Kulshrestha et al. found the inhibitory effect of CaF2-NPs on genes related to S. mutans virulence (gtfC, vicR, comDE, ftf, and spaP). The study also proposed that CaF2-NPs could also suppress the enzymatic activities related to cell adhesion, glucan synthesis, acid production and tolerance, and quorum sensing that result in biofilm inhibition [141]
NPs such as AuNPs (gold NPs) were also developed in combination with hordenine. Hordenine-AuNPs displayed greater antibiofilm properties on P. aeruginosa PAO1, which suggests NPs-delivered natural compounds can be efficiently used in biofilm-related infection [155]. Most recently, Hwang et al., constructed catalytic antimicrobial robots (CARs) that accurately, effectively, and controllably killed, damaged, and detached biofilms [138]. CARs utilizing iron oxide NPs with dual catalytic-magnetic functionality (1) formed free radicals, (2) disrupted biofilm EPS, and (3) eradicated the scrappy biofilm debris using magnetic field-driven robotic assemblies [138].
Nitric oxide-releasing nanoparticles (NO NPs) employ various simultaneous antimicrobial mechanisms, so the chances that microbes will develop resistance to NO NPs is low [149]. Schairer et al. compared hydrogel/glass composite NO NP with systemic vancomycin for treating MRSA-infected intramuscular abscesses in a mouse model and found that NO NPs led to a more significant decrease in bacterial survival than vancomycin treated ones [139]. Drug-delivery NPs with targeting ligands have the potential for promoting improved proximity between the individual biofilm cells within the EPS and the nanocarrier. Lipid and polymer NPs are gaining attraction due to their versatility, biocompatibility, targeted/triggered release, and ability to incorporate lipophilic as well as hydrophilic drugs. Many liposomal formulations for biofilm treatment are under development, yet such products are not available on the market. A liposomal formulation containing the antibiotic amikacin, Arikayce™ (Transave, Inc., Monmouth Junction, NJ, USA) is in Phase III clinical trials, Similarly, Fluidsomes™, containing tobramycin, is in a Phase II trial for the treatment of cystic fibrosis-associated respiratory infections. Although there are some examples of antimicrobial catheters, implants, and wound dressings containing AgNPs available for clinical use, antibiofilm strategies are still underdeveloped. Thus, more in vivo studies for therapeutic applications are required regarding the use of nanoparticles [140].

5.4. Photodynamic Therapy

Photodynamic therapy (PDT) has been utilized for treating several types of viral, fungal, bacterial, protozoan, or even parasitic infections in recent years. It is known that PDT has effectively decreased the clinically important drug-resistant Gram-negative and Gram-positive bacteria [156]. PDT consists of three components: oxygen, visible light, and non-toxic photosensitizers (PS). Since PS-generated ROS can act on several molecules such as DNA, proteins, lipids, biofilm EPS, and even the bacterial cells, PDT appears to have a remarkable possibility in scheming antibiofilm approaches with various targets [157]. Due to its ability of selective binding to the pathogenic cell membranes and particularly targeting the affected tissue for causing extreme harm to microbes while having negligible harm to the host, PDT seems to have some remarkable advantages over conventional methods of treatment [158]. Lately, it was also revealed that PDT could also eliminate the S. mutans biofilms-related infections [12]. Li et al. showed that combining photothermal therapy (PTT) and PDT, aided with glutathione oxidation, offered synergistic rapid killing of S. aureus biofilm bacteria in vivo than that of PTT or PDT alone [159].

6. Future Directions

The detection and extraction of newer compounds are now more convenient with the accessibility of innovative techniques [160]. These new compounds can be developed into effective antimicrobial agents for limiting the formation of microbial biofilms on various surfaces. Furthermore, with a better insight into the biofilm formation process, designing novel therapeutic techniques is much easier. The research on the exploration of novel antimicrobial and antibiofilm agents is ongoing. In this article, we summarized the currently available strategies including the antimicrobial agents and the therapeutic approaches. These techniques could provide new hope to move past the antibiotic era. However, there are more possible prospects in the future. We could also combine multiple approaches to find novel strategies to get rid of persistent biofilms, since biofilms are naturally polymicrobial and resilient [161,162]. However, before using each technique it is necessary to understand the merits and demerits of each agent and select the ones that are most efficient in eliminating targeted biofilms by providing limited damage to the host [163]. For many antibiofilm molecules or tools assessed to date, the main reason for their lack of use in clinical practice is due to the gap between the good outcomes achieved in preclinical studies and the evaluation of their clinical potential [164]. Thus, more clinical trials need to be performed for determining the therapeutic application of the antibiofilm agents [163].
Cancer cells and bacterial cells share some similar traits, such as virulence high replication rates, modes of dispersal within the host, and rapid development of drug resistance [9]. Cancer cells and bacteria share several metabolic features and pathways [9]. Some recent studies revealed the relatio164nship between biofilm flora and cancer development [165,166]. It has also been established that the bacteria used host metabolites to form biofilms and to propagate cancer [165]. There has been some success in repurposing some anticancer drugs such as human kinase inhibitors for biofilm inhibition and eradication including the biofilm persisters [167]. Similarly, Mitomycin C, which is an FDA-approved alkylating agent that is being used as a therapeutic agent for treating many cancers, was also found to be effective in killing the persister cells [9]. Thus, repurposing existing drugs not only appears to be an attractive approach in search of innovative anti-biofilm drugs but also saves time and extensive effort of going through de novo drug optimization processes [167]. Subsequently, the anticancer drugs that target metabolism might play a dual role in targeting cancer as well as bacteria.
Synthetic biology (SB) is a promising interdisciplinary research field that could be used to design and construct newer fabricated devices, artificial metabolic pathways, and organisms, or as well as reform current natural and biological systems, targeting critical problems in health, materials, energy, and the environment [168]. Using several synthetic biological approaches, it is now possible not only to engineer the desired proteins [169] but also to engineer the metabolic pathways [170] or even the biofilm structures [171]. SB also enables us to control biofilms by building quorum-sensing genetic circuits to control biofilms [172,173]. Thus, with better knowledge about biofilms, many novel synthetic biology toolkits could be developed which could expand the potential to control biofilms or even use them for human benefit.

Author Contributions

L.S. and H.-M.F. contributed to the conceptualization, research strategy, and writing, H.-R.T. and J.-D.H. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the National Key Research and Development Program of China (2021YFA0910700, 2018YFA0902701), and Guangdong Science and Technology Department (2020B1212030004) to J.-D.H., Chinese Academy of Sciences ( 2021PB0019) to L.S., Guangdong Provincial Basic and Applied Basic Research Fund Project (2020A1515010726) and Sanming Project of Medicine in Shenzhen (SZSM201911011) to H.-R.T., J.-D.H. is supported by L&T Charitable Foundation and the program for Guangdong Introducing Innovative and Entrepreneurial Teams (2019BT02Y198).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Taku Tokuyasu for his help in critically revising the paper, language editing, and proofreading. The authors also acknowledge financial support from the, National Key Research and Development Program of China (2021YFA0910700, 2018YFA0902701), Guangdong Science and Technology Department (2020B1212030004), Chinese Academy of Sciences President’s International Fellowship Initiative (2021PB0019), Guangdong Provincial Basic and Applied Basic Research Fund Project (2020A1515010726) and Sanming Project of Medicine in Shenzhen (SZSM201911011). J.-D.H. thanks the L&T Charitable Foundation and the Program for Guangdong Introducing Innovative and Entrepreneurial Teams (2019BT02Y198) for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shrestha, L.; Kayama, S.; Sasaki, M.; Kato, F.; Hisatsune, J.; Tsuruda, K.; Koizumi, K.; Tatsukawa, N.; Yu, L.; Takeda, K.; et al. Inhibitory effects of antibiofilm compound 1 against Staphylococcus aureus biofilms. Microbiol. Immunol. 2016, 60, 148–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Watters, C.; Fleming, D.; Bishop, D.; Rumbaugh, K. Chapter seven-host responses to biofilm. Prog. Mol. Biol. Transl. Sci. 2016, 142, 193–239. [Google Scholar]
  3. López, D.; Vlamakis, H.; Kolter, R. Biofilms. Cold Spring Harb. Perspect. Biol. 2010, 2, a000398. [Google Scholar]
  4. Koo, H.; Allan, R.N.; Howlin, R.P.; Stoodley, P.; Hall-Stoodley, L. Targeting microbial biofilms: Current and prospective therapeutic strategies. Nat. Rev. Microbiol. 2017, 15, 740–755. [Google Scholar] [CrossRef] [PubMed]
  5. Bjarnsholt, T. The role of bacterial biofilms in chronic infections. APMIS Suppl. 2013, 136, 1–51. [Google Scholar] [CrossRef]
  6. Hall-Stoodley, L.; Costerton, J.W.; Stoodley, P. Bacterial biofilms: From the natural environment to infectious diseases. Nat. Rev. Microbiol. 2004, 2, 95–108. [Google Scholar] [CrossRef]
  7. Flemming, H.C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563–575. [Google Scholar] [CrossRef]
  8. Jamal, M.; Ahmad, W.; Andleeb, S.; Jalil, F.; Imran, M.; Nawaz, M.A.; Hussain, T.; Ali, M.; Rafiq, M.; Kamil, M.A. Bacterial biofilm and associated infections. J. Chin. Med. Assoc. 2018, 81, 7–11. [Google Scholar] [CrossRef]
  9. Soo, V.; Kwan, B.; Quezada, H.; Castillo-Juárez, I.; Pérez-Eretza, B.; García-Contreras, S.J.; Martínez-Vázquez, M.; Wood, T.K.; García-Contreras, R. Repurposing of anticancer drugs for the treatment of bacterial infections. Curr. Top. Med. Chem. 2016, 17, 1157–1176. [Google Scholar] [CrossRef] [Green Version]
  10. Dincer, S.; Masume Uslu, F.; Delik, A. Antibiotic resistance in biofilm. In Bacterial Biofilms; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef]
  11. Stewart, P.S. Antimicrobial tolerance in bio films. Microbiol. Spectr. 2015, 3, 1–13. [Google Scholar] [CrossRef] [Green Version]
  12. Sharma, D.; Misba, L.; Khan, A.U. Antibiotics versus biofilm: An emerging battleground in microbial communities. Antimicrob. Resist. Infect. Control 2019, 8, 76. [Google Scholar] [CrossRef] [PubMed]
  13. Omar, A.; Wright, J.B.; Schultz, G.; Burrell, R.; Nadworny, P. Microbial biofilms and chronic wounds. Microorganisms 2017, 5, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Davies, D. Understanding biofilm resistance to antibacterial agents. Nat. Rev. Drug Discov. 2003, 2, 114–122. [Google Scholar] [CrossRef] [PubMed]
  15. Pogodin, S.; Hasan, J.; Baulin, V.A.; Webb, H.K.; Truong, V.K.; Phong Nguyen, T.H.; Boshkovikj, V.; Fluke, C.J.; Watson, G.S.; Watson, J.A.; et al. Biophysical model of bacterial cell interactions with nanopatterned cicada wing surfaces. Biophys. J. 2013, 104, 835–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Tang, P.; Zhang, W.; Wang, Y.; Zhang, B.; Wang, H.; Lin, C.; Zhang, L. Effect of superhydrophobic surface of titanium on Staphylococcus aureus adhesion. J. Nanomater. 2011, 2011, 178921. [Google Scholar] [CrossRef] [Green Version]
  17. Kim, S.; Jung, U.T.; Kim, S.K.; Lee, J.H.; Choi, H.S.; Kim, C.S.; Jeong, M.Y. Nanostructured multifunctional surface with antireflective and antimicrobial characteristics. ACS Appl. Mater. Interfaces 2015, 7, 326–331. [Google Scholar] [CrossRef]
  18. Pereira, M.O.; Machado, I.; Simoes, M.; Viera, M.J. Preventing biofilm formation using surfactants. Biofilm Club 2007, 167–174. [Google Scholar]
  19. Brackman, G.; Coenye, T. Quorum sensing inhibitors as anti-biofilm agents. Curr. Pharm. Des. 2014, 21, 5–11. [Google Scholar] [CrossRef]
  20. Lijuan, C.; Xing, Y.; Minxi, W.; Wenkai, L.; Le, D. Development of an aptamer-ampicillin conjugate for treating biofilms. Biochem. Biophys. Res. Commun. 2017, 483, 847–854. [Google Scholar] [CrossRef]
  21. Yasuda, H.; Ajiki, Y.; Koga, T.; Kawada, H.; Yokota, T. Interaction between biofilms formed by Pseudomonas aeruginosa and clarithromycin. Antimicrob. Agents Chemother. 1993, 37, 1749–1755. [Google Scholar] [CrossRef] [Green Version]
  22. Kim, C.; Kim, J.; Park, H.Y.; Lee, J.H.; Park, H.J.; Kim, C.K.; Yoon, J. Structural understanding of quorum-sensing inhibitors by molecular modeling study in Pseudomonas aeruginosa. Appl. Microbiol. Biotechnol. 2009, 83, 1095–1103. [Google Scholar] [CrossRef] [PubMed]
  23. Kim, C.; Kim, J.; Park, H.Y.; Park, H.J.; Kim, C.K.; Yoon, J.; Lee, J.H. Development of inhibitors against TraR quorum-sensing system in Agrobacterium tumefaciens by molecular modeling of the ligand-receptor interaction. Mol. Cells 2009, 28, 447–453. [Google Scholar] [CrossRef] [PubMed]
  24. Ren, D.; Sims, J.J.; Wood, T.K. Inhibition of biofilm formation and swarming of Escherichia coli by (5Z)-4-bromo-5-(bromomethylene)-3- butyl-2(5H)-furanone. Environ. Microbiol. 2001, 3, 731–736. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, C.; Kim, J.; Park, H.Y.; Park, H.J.; Lee, J.H.; Kim, C.K.; Yoon, J. Furanone derivatives as quorum-sensing antagonists of Pseudomonas aeruginosa. Appl. Microbiol. Biotechnol. 2008, 80, 37–47. [Google Scholar] [CrossRef]
  26. Kim, J.; Pitts, B.; Stewart, P.S.; Camper, A.; Yoon, J. Comparison of the antimicrobial effects of chlorine, silver ion, and tobramycin on biofilm. Antimicrob. Agents Chemother. 2008, 52, 1446–1453. [Google Scholar] [CrossRef] [Green Version]
  27. Li, X.H.; Lee, J.H. Antibiofilm agents: A new perspective for antimicrobial strategy. J. Microbiol. 2017, 55, 753–766. [Google Scholar] [CrossRef]
  28. Plakunov, V.K.; Zhurina, M.V.; Gannesen, A.V.; Mart’yanov, S.V.; Nikolaev, Y.A. Antibiofilm agents: Therminological ambiguity and strategy for search. Microbiology 2019, 88, 747–750. [Google Scholar] [CrossRef]
  29. Khalid, S.; Gao, A.; Wang, G.; Chu, P.K.; Wang, H. Tuning surface topographies on biomaterials to control bacterial infection. Biomater. Sci. 2020, 8, 6840–6857. [Google Scholar] [CrossRef]
  30. Allegrone, G.; Ceresa, C.; Rinaldi, M.; Fracchia, L. Diverse effects of natural and synthetic surfactants on the inhibition of staphylococcus aureus biofilm. Pharmaceutics 2021, 13, 1172. [Google Scholar] [CrossRef]
  31. Ohta, K.; Komatsuzawa, H.; Sugai, M.; Suginaka, H. Triton X-100-induced lipoteichoic acid release is correlated with the methicillin resistance in Staphylococcus aureus. FEMS Microbiol. Lett. 2000, 182, 77–79. [Google Scholar] [CrossRef]
  32. Komatsuzawa, H.; Suzuki, J.; Sugai, M.; Miyake, Y.; Suginaka, H. The effect of triton X-100 on the in-vitro susceptibility of methicillin-resistant Staphylococcus aureus to oxacillin. J. Antimicrob. Chemother. 1994, 34, 3537–3545. [Google Scholar] [CrossRef] [PubMed]
  33. Nitschke, M.; Costa, S.G.V.A.O. Biosurfactants in food industry. Trends Food Sci. Technol. 2007, 18, 252–259. [Google Scholar] [CrossRef]
  34. Percival, S.L.; Mayer, D.; Kirsner, R.S.; Schultz, G.; Weir, D.; Roy, S.; Alavi, A.; Romanelli, M. Surfactants: Role in biofilm management and cellular behaviour. Int. Wound J. 2019, 16, 753–760. [Google Scholar] [CrossRef] [PubMed]
  35. Abdel-Mawgoud, A.M.; Lépine, F.; Déziel, E. Rhamnolipids: Diversity of structures, microbial origins and roles. Appl. Microbiol. Biotechnol. 2010, 86, 1323–1336. [Google Scholar] [CrossRef] [Green Version]
  36. ESilva, S.S.; Carvalho, J.W.P.; Aires, C.P.; Nitschke, M. Disruption of Staphylococcus aureus biofilms using rhamnolipid biosurfactants. J. Dairy Sci. 2017, 100, 7864–7873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Chebbi, A.; Elshikh, M.; Haque, F.; Ahmed, S.; Dobbin, S.; Marchant, R.; Sayadi, S.; Chamkha, M.; Banat, I.M. Rhamnolipids from Pseudomonas aeruginosa strain W10; as antibiofilm/antibiofouling products for metal protection. J. Basic Microbiol. 2017, 57, 364–375. [Google Scholar] [CrossRef]
  38. Banat, I.M.; de Rienzo, M.A.D.; Quinn, G.A. Microbial biofilms: Biosurfactants as antibiofilm agents. Appl. Microbiol. Biotechnol. 2014, 98, 9915–9929. [Google Scholar] [CrossRef]
  39. D’Auria, L.; Deleu, M.; Dufour, S.; Mingeot-Leclercq, M.P.; Tyteca, D. Surfactins modulate the lateral organization of fluorescent membrane polar lipids: A new tool to study drug: Membrane interaction and assessment of the role of cholesterol and drug acyl chain length. Biochim. Biophys. Acta-Biomembr. 2013, 1828, 2064–2073. [Google Scholar] [CrossRef] [Green Version]
  40. Finnegan, S.; Percival, S.L. EDTA: An antimicrobial and antibiofilm agent for use in wound care. Adv. Wound Care 2015, 4, 415–421. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, A.; Mu, H.; Zhang, W.; Cui, G.; Zhu, J.; Duan, J. Chitosan coupling makes microbial biofilms susceptible to antibiotics. Sci. Rep. 2013, 3, 3364. [Google Scholar] [CrossRef] [Green Version]
  42. Vikram, A.; Jesudhasan, P.R.; Jayaprakasha, G.K.; Pillai, S.D.; Patil, B.S. Citrus limonoids interfere with Vibrio harveyi cell–cell signalling and biofilm formation by modulating the response regulator luxO. Microbiology 2011, 157, 99–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Scopel, M.; Abraham, W.R.; Henriques, A.T.; MacEdo, A.J. Dipeptide cis-cyclo(Leucyl-Tyrosyl) produced by sponge associated Penicillium sp. F37 inhibits biofilm formation of the pathogenic Staphylococcus epidermidis. Bioorganic Med. Chem. Lett. 2013, 23, 624–626. [Google Scholar] [CrossRef] [PubMed]
  44. Park, S.R.; Tripathi, A.; Wu, J.; Schultz, P.J.; Yim, I.; McQuade, T.J.; Yu, F.; Arevang, C.J.; Mensah, A.Y.; Castillo, G.T.; et al. Discovery of cahuitamycins as biofilm inhibitors derived from a convergent biosynthetic pathway. Nat. Commun. 2016, 7, 10710. [Google Scholar] [CrossRef] [PubMed]
  45. Eom, S.H.; Lee, D.S.; Jung, Y.J.; Park, J.H.; Choi, J.I.; Yim, M.J.; Jeon, J.M.; Kim, H.W.; Son, K.T.; Je, J.Y.; et al. The mechanism of antibacterial activity of phlorofucofuroeckol-A against methicillin-resistant Staphylococcus aureus. Appl. Microbiol. Biotechnol. 2014, 98, 504–518. [Google Scholar] [CrossRef] [PubMed]
  46. Kalpana, B.J.; Aarthy, S.; Pandian, S.K. Antibiofilm activity of α-amylase from Bacillus subtilis S8-18 against biofilm forming human bacterial pathogens. Appl. Biochem. Biotechnol. 2012, 167, 1778–1794. [Google Scholar] [CrossRef]
  47. Kolodkin-Gal, I.; Cao, S.; Chai, L.; Böttcher, T.; Kolter, R.; Clardy, J.; Losick, R. A self-produced trigger for biofilm disassembly that targets exopolysaccharide. Cell 2012, 149, 684–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kolodkin-Gal, I.; Romero, D.; Cao, S.; Clardy, J.; Kolter, R.; Losick, R. D-Amino acids trigger biofilm disassembly. Science 2010, 328, 627–629. [Google Scholar] [CrossRef] [Green Version]
  49. Muñoz-Egea, M.C.; García-Pedrazuela, M.; Mahillo-Fernandez, I.; Esteban, J. Effect of antibiotics and antibiofilm agents in the ultrastructure and development of biofilms developed by nonpigmented rapidly growing mycobacteria. Microb. Drug Resist. 2016, 22, 1–6. [Google Scholar] [CrossRef]
  50. Iwase, T.; Uehara, Y.; Shinji, H.; Tajima, A.; Seo, H.; Takada, K.; Agata, T.; Mizunoe, Y. Staphylococcus epidermidis Esp inhibits Staphylococcus aureus biofilm formation and nasal colonization. Nature 2010, 465, 346–349. [Google Scholar] [CrossRef]
  51. Tetz, G.V.; Artemenko, N.K.; Tetz, V.V. Effect of DNase and antibiotics on biofilm characteristics. Antimicrobial Agents Chemother. 2009, 53, 1204–1209. [Google Scholar] [CrossRef] [Green Version]
  52. Zhao, X.; Liu, Z.Z.; Liu, Z.Z.; Meng, R.; Shi, C.; Chen, X.; Bu, X.; Guo, N. Phenotype and RNA-seq-based transcriptome profiling of Staphylococcus aureus biofilms in response to tea tree oil. Microb. Pathog. 2018, 123, 304–313. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, H.; Moser, C.; Wang, H.Z.; Høiby, N.; Song, Z.J. Strategies for combating bacterial biofilm infections. Int. J. Oral Sci. 2015, 7, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Roy, R.; Tiwari, M.; Donelli, G.; Tiwari, V. Strategies for combating bacterial biofilms: A focus on anti-biofilm agents and their mechanisms of action. Virulence 2018, 9, 522–554. [Google Scholar] [CrossRef] [PubMed]
  55. Hoeltje, J.v.; Mirelman, D.; Sharon, N.; Schwarz, U. Novel type of murein transglycosylase in Escherichia coli. J. Bacteriol. 1975, 124, 1067–1076. [Google Scholar] [CrossRef] [Green Version]
  56. Shen, Y.; Köller, T.; Kreikemeyer, B.; Nelson, D.C. Rapid degradation of Streptococcus pyogenes biofilms by PlyC, a bacteriophage-encoded endolysin. J. Antimicrob. Chemother. 2013, 8, 1818–1824. [Google Scholar] [CrossRef] [PubMed]
  57. Worthington, R.J.; Richards, J.J.; Melander, C. Small molecule control of bacterial biofilms. Org. Biomol. Chem. 2012, 10, 7457–7474. [Google Scholar] [CrossRef] [Green Version]
  58. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199. [Google Scholar] [CrossRef] [Green Version]
  59. Romling, U.; Galperin, M.Y.; Gomelsky, M. Cyclic di-GMP: The first 25 years of a universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 2013, 77, 1–52. [Google Scholar] [CrossRef] [Green Version]
  60. 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] [Green Version]
  61. Ryan, R.P.; Fouhy, Y.; Lucey, J.F.; Dow, J.M. Cyclic di-GMP signaling in bacteria: Recent advances and new puzzles. J. Bacteriol. 2006, 188, 8327–8334. [Google Scholar] [CrossRef] [Green Version]
  62. Harjai, K.; Kumar, R.; Singh, S. Garlic blocks quorum sensing and attenuates the virulence of Pseudomonas aeruginosa. FEMS Immunol. Med. Microbiol. 2010, 58, 161–168. [Google Scholar] [CrossRef] [Green Version]
  63. Persson, T.; Hansen, T.H.; Rasmussen, T.B.; Skindersø, M.E.; Givskov, M.; Nielsen, J. Rational design and synthesis of new quorum-sensing inhibitors derived from acylated homoserine lactones and natural products from garlic. Org. Biomol. Chem. 2005, 3, 253–262. [Google Scholar] [CrossRef] [PubMed]
  64. Vikram, A.; Jesudhasan, P.R.; Pillai, S.D.; Patil, B.S. Isolimonic acid interferes with Escherichia coli O157, H7 biofilm and TTSS in QseBC and QseA dependent fashion. BMC Microbiol. 2012, 12, 261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Brackman, G.; Defoirdt, T.; Miyamoto, C.; Bossier, P.; Calenbergh, S.; van Nelis, H.; Coenye, T. Cinnamaldehyde and cinnamaldehyde derivatives reduce virulence in Vibrio spp. by decreasing the DNA-binding activity of the quorum sensing response regulator LuxR. BMC Microbiol. 2008, 8, 149. [Google Scholar] [CrossRef] [Green Version]
  66. Niu, C.; Gilbert, E.S. Colorimetric method for identifying plant essential oil components that affect biofilm formation and structure. Appl. Environ. Microbiol. 2004, 70, 9651–9656. [Google Scholar] [CrossRef] [Green Version]
  67. Zhou, J.W.; Luo, H.Z.; Jiang, H.; Jian, T.K.; Chen, Z.Q.; Jia, A.Q. Hordenine: A novel quorum sensing inhibitor and antibiofilm agent against Pseudomonas aeruginosa. J. Agric. Food Chem. 2018, 66, 1620–1628. [Google Scholar] [CrossRef] [PubMed]
  68. Zhou, J.W.; Hou, B.; Liu, G.Y.; Jiang, H.; Sun, B.; Wang, Z.N.; Shi, R.F.; Xu, Y.; Wang, R.; Jia, A.Q. Attenuation of Pseudomonas aeruginosa biofilm by hordenine: A combinatorial study with aminoglycoside antibiotics. Appl. Microbiol. Biotechnol. 2018, 102, 9745–9758. [Google Scholar] [CrossRef] [PubMed]
  69. Krishnan, T.; Yin, W.F.; Chan, K.G. Inhibition of quorum sensing-controlled virulence factor production in Pseudomonas aeruginosa PAO1 by ayurveda spice clove (syzygium aromaticum) bud extract. Sensors 2012, 12, 4016–4030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Tan, S.Y.Y.; Chua, S.L.; Chen, Y.; Rice, S.A.; Kjelleberg, S.; Nielsen, T.E.; Yang, L.; Givskov, M. Identification of five structurally unrelated quorum-sensing inhibitors of Pseudomonas aeruginosa from a natural-derivative database. Antimicrob. Agents Chemother. 2013, 57, 5629–5641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Chu, W.; Zhou, S.; Jiang, Y.; Zhu, W.; Zhuang, X.; Fu, J. Effect of traditional Chinese herbal medicine with antiquorum sensing activity on Pseudomonas aeruginosa. Evid.-Based Complementary Altern. Med. 2013, 2013, 648257. [Google Scholar] [CrossRef] [Green Version]
  72. Vasavi, H.S.; Arun, A.B.; Rekha, P.D. Anti-quorum sensing activity of flavonoid-rich fraction from Centella asiatica L. against Pseudomonas aeruginosa PAO1. J. Microbiol. Immunol. Infect. 2016, 49, 8–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ouyang, J.; Sun, F.; Feng, W.; Sun, Y.; Qiu, X.; Xiong, L.; Liu, Y.; Chen, Y. Quercetin is an effective inhibitor of quorum sensing, biofilm formation and virulence factors in Pseudomonas aeruginosa. J. Appl. Microbiol. 2016, 120, 966–974. [Google Scholar] [CrossRef] [Green Version]
  74. Lauderdale, K.J.; Malone, C.L.; Boles, B.R.; Morcuende, J.; Horswill, A.R. Biofilm dispersal of community-associated methicillin-resistant Staphylococcus aureus on orthopedic implant material. J. Orthop. Res. 2010, 28, 55–61. [Google Scholar] [CrossRef]
  75. Simonetti, O.; Cirioni, O.; Ghiselli, R.; Goteri, G.; Scalise, A.; Orlando, F.; Silvestri, C.; Riva, A.; Saba, V.; Madanahally, K.D.; et al. RNAIII-inhibiting peptide enhances healing of wounds infected with methicillin-resistant Staphylococcus aureus. Antimicrob. Agents Chemother. 2008, 52, 2205–2211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Barraud, N.; Schleheck, D.; Klebensberger, J.; Webb, J.S.; Hassett, D.J.; Rice, S.A.; Kjelleberg, S. Nitric oxide signaling in Pseudomonas aeruginosa biofilms mediates phosphodiesterase activity, decreased cyclic di-GMP levels, and enhanced dispersal. J. Bacteriol. 2009, 191, 7333–7342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Schreiber, F.; Beutler, M.; Enning, D.; Lamprecht-Grandio, M.; Zafra, O.; González-Pastor, J.; de Beer, D. The role of nitric-oxide-synthase-derived nitric oxide in multicellular traits of Bacillus subtilis 3610, Biofilm formation, swarming, and dispersal. BMC Microbiol. 2011, 11, 111. [Google Scholar] [CrossRef] [Green Version]
  78. Parrino, B.; Schillaci, D.; Carnevale, I.; Giovannetti, E.; Diana, P.; Cirrincione, G.; Cascioferro, S. Synthetic small molecules as anti-biofilm agents in the struggle against antibiotic resistance. Eur. J. Med. Chem. 2019, 161, 154–178. [Google Scholar] [CrossRef] [PubMed]
  79. Robijns, S.C.A.; Roberfroid, S.; van Puyvelde, S.; de Pauw, B.; Uceda Santamaría, E.; de Weerdt, A.; De Coster, D.; Hermans, K.; De Keersmaecker, S.C.; Vanderleyden, J.; et al. A GFP promoter fusion library for the study of Salmonella biofilm formation and the mode of action of biofilm inhibitors. Biofouling 2014, 30, 605–625. [Google Scholar] [CrossRef] [PubMed]
  80. Sambanthamoorthy, K.; Gokhale, A.A.; Lao, W.; Parashar, V.; Neiditch, M.B.; Semmelhack, M.F.; Lee, I.; Waters, C.M. Identification of a novel benzimidazole that inhibits bacterial biofilm formation in a broad-spectrum manner. Antimicrob. Agents Chemother. 2011, 55, 4369–4378. [Google Scholar] [CrossRef] [Green Version]
  81. Coldham, N.G.; Webber, M.; Woodward, M.J.; Piddock, L.J.V. A 96-well plate fluorescence assay for assessment of cellular permeability and active efflux in Salmonella enterica serovar Typhimurium and Escherichia coli. J. Antimicrob. Chemother. 2010, 65, 1655–1663. [Google Scholar] [CrossRef] [Green Version]
  82. Suresh, L.; Sagar Vijay Kumar, P.; Poornachandra, Y.; Ganesh Kumar, C.; Chandramouli, G.V.P. Design, synthesis and evaluation of novel pyrazolo-pyrimido [4,5-d] pyrimidine derivatives as potent antibacterial and biofilm inhibitors. Bioorganic Med. Chem. Lett. 2017, 27, 1451–1457. [Google Scholar] [CrossRef] [PubMed]
  83. Jitender Dev, G.; Poornachandra, Y.; Ratnakar Reddy, K.; Naresh Kumar, R.; Ravikumar, N.; Krishna Swaroop, D.; Ranjithreddy, P.; Shravan Kumar, G.; Nanubolu, J.B.; Ganesh Kumar, C.; et al. Synthesis of novel pyrazolo [3,4-b] quinolinyl acetamide analogs, their evaluation for antimicrobial and anticancer activities, validation by molecular modeling and CoMFA analysis. Eur. J. Med. Chem. 2017, 130, 223–239. [Google Scholar] [CrossRef]
  84. Dixon, S.J.; Stockwell, B.R. The role of iron and reactive oxygen species in cell death. Nat. Chem. Biol. 2014, 10, 9–17. [Google Scholar] [CrossRef] [PubMed]
  85. Ryu, J.H.; Beuchat, L.R. Biofilm formation by Escherichia coli O157, H7 on stainless steel: Effect of exopolysaccharide and curli production on its resistance to chlorine. Appl. Environ. Microbiol. 2005, 71, 247–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Uhlich, G.A.; Cooke, P.H.; Solomon, E.B. Analyses of the red-dry-rough phenotype of an Escherichia coli O157, H7 strain and its role in biofilm formation and resistance to antibacterial agents. Appl. Environ. Microbiol. 2006, 72, 2564–2572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Maggio, B.; Raffa, D.; Raimondi, M.V.; Cascioferro, S.; Plescia, F.; Schillaci, D.; Cusimano, M.G.; Leonchiks, A.; Zhulenkovs, D.; Basile, L.; et al. Discovery of a new class of sortase a transpeptidase inhibitors to tackle gram-positive pathogens: 2-(2-phenylhydrazinylidene)alkanoic acids and related derivatives. Molecules 2016, 21, 241. [Google Scholar] [CrossRef] [Green Version]
  88. Mazmanian, S.K.; Liu, G.; Jensen, E.R.; Lenoy, E.; Schneewind, O. Staphylococcus aureus sortase mutants defective in the display of surface proteins and in the pathogenesis of animal infections. Proc. Natl. Acad. Sci. USA 2000, 97, 5510–5515. [Google Scholar] [CrossRef] [Green Version]
  89. Oh, K.B.; Oh, M.N.; Kim, J.G.; Shin, D.S.; Shin, J. Inhibition of sortase-mediated Staphylococcus aureus adhesion to fibronectin via fibronectin-binding protein by sortase inhibitors. Appl. Microbiol. Biotechnol. 2006, 70, 102–106. [Google Scholar] [CrossRef]
  90. Donlan, R.M.; Costerton, J.W. Biofilms: Survival mechanisms of clinically relevant microorganisms. Clin. Microbiol. Rev. 2002, 15, 167–193. [Google Scholar] [CrossRef] [Green Version]
  91. Lönn-Stensrud, J.; Petersen, F.C.; Benneche, T.; Scheie, A.A. Synthetic bromated furanone inhibits autoinducer-2-mediated communication and biofilm formation in oral streptococci. Oral Microbiol. Immunol. 2007, 22, 340–346. [Google Scholar] [CrossRef]
  92. Garrison, A.T.; Abouelhassan, Y.; Norwood, V.M.; Kallifidas, D.; Bai, F.; Nguyen, M.T.; Rolfe, M.; Burch, G.M.; Jin, S.; Luesch, H.; et al. Structure-activity relationships of a diverse class of halogenated phenazines that targets persistent, antibiotic-tolerant bacterial biofilms and Mycobacterium tuberculosis. J. Med. Chem. 2016, 59, 3808–3825. [Google Scholar] [CrossRef] [PubMed]
  93. Soukarieh, F.; Oton, E.V.; Dubern, J.F.; Gomes, J.; Halliday, N.; De Pilar Crespo, M.; Prada, J.R.; Insuasty, B.; Abonia, R.; Quiroga, J.; et al. In silico and in vitro-guided identification of inhibitors of alkylquinolone-dependent quorum sensing in Pseudomonas aeruginosa. Molecules 2018, 23, 257. [Google Scholar] [CrossRef] [Green Version]
  94. Sommer, R.; Hauck, D.; Varrot, A.; Wagner, S.; Audfray, A.; Prestel, A.; Moller, H.M.; Imberty, A.; Titz, A. Cinnamide Derivatives of d-mannose as inhibitors of the bacterial virulence factor lecb from Pseudomonas aeruginosa. ChemistryOpen 2015, 4, 756–767. [Google Scholar] [CrossRef]
  95. Petersen, L.-E.; Kellermann, M.Y.; Schupp, P.J. Secondary metabolites of marine microbes: From natural products chemistry to chemical ecology. YOUMARES 2020, 9, 159–180. [Google Scholar] [CrossRef] [Green Version]
  96. Adnan, M.; Alshammari, E.; Patel, M.; Ashraf, S.A.; Khan, S.; Hadi, S. Significance and potential of marine microbial natural bioactive compounds against biofilms/biofouling: Necessity for green chemistry. PeerJ 2018, 2018, e5049. [Google Scholar] [CrossRef] [PubMed]
  97. Rosa, G.P.; Tavares, W.R.; Sousa, P.M.C.; Pagès, A.K.; Seca, A.M.L.; Pinto, D.C.G.A. Seaweed secondary metabolites with beneficial health effects: An overview of successes in in vivo studies and clinical trials. Mar. Drugs 2020, 18, 8. [Google Scholar] [CrossRef] [Green Version]
  98. Wang, Z.; Shen, Y.; Haapasalo, M. Antibiofilm peptides against oral biofilms. J. Oral Microbiol. 2017, 9, 1327308. [Google Scholar] [CrossRef] [Green Version]
  99. Gaspar, D.; Veiga, A.S.; Castanho, M.A.R.B. From antimicrobial to anticancer peptides. A review. Front. Microbiol. 2013, 4, 294. [Google Scholar] [CrossRef] [Green Version]
  100. Giuliani, A.; Pirri, G.; Nicoletto, S.F. Antimicrobial peptides: An overview of a promising class of therapeutics. Cent. Eur. J. Biol. 2007, 2, 1–33. [Google Scholar] [CrossRef]
  101. Mohamed, M.F.; Abdelkhalek, A.; Seleem, M.N. Evaluation of short synthetic antimicrobial peptides for treatment of drug-resistant and intracellular Staphylococcus aureus. Sci. Rep. 2016, 6, 29707. [Google Scholar] [CrossRef]
  102. De La Fuente-Núñez, C.; Reffuveille, F.; Haney, E.F.; Straus, S.K.; Hancock, R.E.W. Broad-spectrum anti-biofilm peptide that targets a cellular stress response. PLoS Pathog. 2014, 10, e1004152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Andrea, A.; Molchanova, N.; Jenssen, H. Antibiofilm peptides and peptidomimetics with focus on surface immobilization. Biomolecules 2018, 8, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Pimentel-Filho, N.d.J.; Martins, M.C.d.F.; Nogueira, G.B.; Mantovani, H.C.; Vanetti, M.C.D. Bovicin HC5 and nisin reduce Staphylococcus aureus adhesion to polystyrene and change the hydrophobicity profile and Gibbs free energy of adhesion. Int. J. Food Microbiol. 2014, 190, 1–8. [Google Scholar] [CrossRef] [Green Version]
  105. Overhage, J.; Campisano, A.; Bains, M.; Torfs, E.C.W.; Rehm, B.H.A.; Hancock, R.E.W. Human host defense peptide LL-37 prevents bacterial biofilm formation. Infect. Immun. 2008, 76, 4176–4182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. De La Fuente-Núñez, C.; Korolik, V.; Bains, M.; Nguyen, U.; Breidenstein, E.B.; Horsman, S.; Lewenza, S.; Burrows, L.; Hancock, R.E. Inhibition of bacterial biofilm formation and swarming motility by a small synthetic cationic peptide. Antimicrob. Agents Chemother. 2012, 56, 2696–2704. [Google Scholar] [CrossRef] [Green Version]
  107. De La Fuente-Núñez, C.; Reffuveille, F.; Mansour, S.C.; Reckseidler-Zenteno, S.L.; Hernández, D.; Brackman, G.; Coenye, T.; Hancock, E.W. D-enantiomeric peptides that eradicate wild-type and multidrug-resistant biofilms and protect against lethal Pseudomonas aeruginosa infections. Chem. Biol. 2015, 22, 196–205. [Google Scholar] [CrossRef]
  108. Otvos, L.; Insug, O.; Rogers, M.E.; Consolvo, P.J.; Condie, B.A.; Lovas, S.; Bulet, P.; Blaszczyk-Thurin, M. Interaction between heat shock proteins and antimicrobial peptides. Biochemistry 2000, 39, 14150–14159. [Google Scholar] [CrossRef]
  109. Kragol, G.; Lovas, S.; Varadi, G.; Condie, B.A.; Hoffmann, R.; Otvos, L. The antibacterial peptide pyrrhocoricin inhibits the ATPase actions of DnaK and prevents chaperone-assisted protein folding. Biochemistry 2001, 40, 3016–3026. [Google Scholar] [CrossRef]
  110. Kragol, G.; Hoffmann, R.; Chattergoon, M.A.; Lovas, S.; Cudic, M.; Bulet, P.; Condie, B.A.; Rosengren, K.J.; Montaner, L.J.; Otvos, L., Jr. Identification of crucial residues for the antibacterial activity of the proline-rich peptide, pyrrhocoricin. Eur. J. Biochem. 2002, 269, 4226–4237. [Google Scholar] [CrossRef]
  111. Vizan, J.L.; Hernandez-Chico, C.; Del Castillo, I.; Moreno, F. The peptide antibiotic microcin B17 induces double-strand cleavage of DNA mediated by E. coli DNA gyrase. EMBO J. 1991, 10, 467–476. [Google Scholar] [CrossRef]
  112. Boman, H.G.; Agerberth, B.; Boman, A. Mechanisms of action on Escherichia coli of cecropin P1 and PR-39, two antibacterial peptides from pig intestine. Infect. Immun. 1993, 61, 2978–2984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Luca, V.; Stringaro, A.; Colone, M.; Pini, A.; Mangoni, M.L. Esculentin(1–21), an amphibian skin membrane-active peptide with potent activity on both planktonic and biofilm cells of the bacterial pathogen Pseudomonas aeruginosa. Cell. Mol. Life Sci. 2013, 70, 2773–2786. [Google Scholar] [CrossRef] [PubMed]
  114. Pulido, D.; Prats-Ejarque, G.; Villalba, C.; Albacar, M.; González-López, J.J.; Torrent, M.; Moussaoui, M.; Boix, E. A novel RNase 3/ECP peptide for Pseudomonas aeruginosa biofilm eradication that combines antimicrobial, lipopolysaccharide binding, and cell-agglutinating activities. Antimicrob. Agents Chemother. 2016, 60, 6313–6325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Quilès, F.; Saadi, S.; Francius, G.; Bacharouche, J.; Humbert, F. In situ and real time investigation of the evolution of a Pseudomonas fluorescens nascent biofilm in the presence of an antimicrobial peptide. Biochim. Biophys. Acta-Biomembr. 2016, 1858, 75–84. [Google Scholar] [CrossRef] [PubMed]
  116. Okuda, K.I.; Zendo, T.; Sugimoto, S.; Iwase, T.; Tajima, A.; Yamada, S.; Sonomoto, K.; Mizunoe, Y. Effects of bacteriocins on methicillin-resistant Staphylococcus aureus biofilm. Antimicrob. Agents Chemother. 2013, 57, 5572–5579. [Google Scholar] [CrossRef] [Green Version]
  117. Brancatisano, F.L.; Maisetta, G.; di Luca, M.; Esin, S.; Bottai, D.; Bizzarri, R.; Campa, M.; Batoni, G. Inhibitory effect of the human liver-derived antimicrobial peptide hepcidin 20 on biofilms of polysaccharide intercellular adhesin (PIA)-positive and PIA-negative strains of Staphylococcus epidermidis. Biofouling 2014, 30, 435–446. [Google Scholar] [CrossRef]
  118. Zhu, C.; Tan, H.; Cheng, T.; Shen, H.; Shao, J.; Guo, Y.; Shi, S.; Zhang, X. Human β-defensin 3 inhibits antibiotic-resistant Staphylococcus biofilm formation. J. Surg. Res. 2013, 183, 204–213. [Google Scholar] [CrossRef]
  119. Ansari, J.M.; Abraham, N.M.; Massaro, J.; Murphy, K.; Smith-Carpenter, J.; Fikrig, E. Anti-biofilm activity of a self-aggregating peptide against Streptococcus mutans. Front. Microbiol. 2017, 8, 488. [Google Scholar] [CrossRef]
  120. Hancock, R.E.W.; Alford, M.A.; Haney, E.F. Antibiofilm activity of host defence peptides: Complexity provides opportunities. Nat. Rev. Microbiol. 2021, 19, 786–797. [Google Scholar] [CrossRef]
  121. Allison, K.R.; Brynildsen, M.P.; Collins, J.J. Metabolite-enabled eradication of bacterial persisters by aminoglycosides. Nature 2011, 473, 216–220. [Google Scholar] [CrossRef] [Green Version]
  122. Pisithkul, T.; Schroeder, J.W.; Trujillo, E.A.; Yeesin, P.; Stevenson, D.M.; Chaiamarit, T.; Coon, J.J.; Wang, J.D.; Amador-Noguez, D. Metabolic remodeling during biofilm development of bacillus subtilis. MBio 2019, 10, e00623-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Lu, H.; Que, Y.; Wu, X.; Guan, T.; Guo, H. Metabolomics deciphered metabolic reprogramming required for biofilm formation. Sci. Rep. 2019, 9, 13160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. He, J.; Hwang, G.; Liu, Y.; Gao, L.; Kilpatrick-Liverman, L.T.; Santarpi, P.; Zhou, X.; Koo, H. L-arginine modifies the exopolysaccharide matrix and thwarts Streptococcus mutans outgrowth within mixed-species oral biofilms. J. Bacteriol. 2016, 198, 2651–2661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Sun, F.; Qu, F.; Ling, Y.; Mao, P.; Xia, P.; Chen, H.; Zhou, D. Biofilm-associated infections: Antibiotic resistance and novel therapeutic strategies. Future Microbiol. 2013, 8, 877–886. [Google Scholar] [CrossRef] [PubMed]
  126. Hochbaum, A.I.; Kolodkin-Gal, I.; Foulston, L.; Kolter, R.; Aizenberg, J.; Losick, R. Inhibitory effects of D-amino acids on Staphylococcus aureus biofilm development. J. Bacteriol. 2011, 193, 5616–5622. [Google Scholar] [CrossRef] [Green Version]
  127. Furfaro, L.L.; Payne, M.S.; Chang, B.J. Bacteriophage therapy: Clinical trials and regulatory hurdles. Front. Cell. Infect. Microbiol. 2018, 8, 376. [Google Scholar] [CrossRef] [Green Version]
  128. Tkhilaishvili, T.; Lombardi, L.; Klatt, A.B.; Trampuz, A.; Di Luca, M. Bacteriophage Sb-1 enhances antibiotic activity against biofilm, degrades exopolysaccharide matrix and targets persisters of Staphylococcus aureus. Int. J. Antimicrob. Agents 2018, 52, 842–853. [Google Scholar] [CrossRef]
  129. Harper, D.R.; Parracho, H.M.R.T.; Walker, J.; Sharp, R.; Hughes, G.; Werthén, M.; Lehman, S.; Morales, S. Bacteriophages and biofilms. Antibiotics 2014, 3, 270–284. [Google Scholar] [CrossRef]
  130. Abedon, S.T. Bacteriophage exploitation of bacterial biofilms: Phage preference for less mature targets? FEMS Microbiol. Lett. 2016, 363, fnv246. [Google Scholar] [CrossRef]
  131. Gou, Y.; Liu, W.; Wang, J.J.; Tan, L.; Hong, B.; Guo, L.; Liu, H.; Pan, Y.; Zhao, Y. Crispr-cas9 knockout of qseb induced asynchrony between motility and biofilm formation in escherichia coli. Can. J. Microbiol. 2019, 65, 691–702. [Google Scholar] [CrossRef]
  132. Zuberi, A.; Misba, L.; Khan, A.U. CRISPR interference (CRISPRi) inhibition of luxS gene expression in E. coli: An approach to inhibit biofilm. Front. Cell. Infect. Microbiol. 2017, 7, 214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Zuberi, A.; Ahmad, N.; Khan, A.U. CRISPRi induced suppression of fimbriae gene (fimH) of a uropathogenic Escherichia coli: An approach to inhibit microbial biofilms. Front. Immunol. 2017, 8, 1552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hegde, S.; Nilyanimit, P.; Kozlova, E.; Anderson, E.R.; Narra, H.P.; Sahni, S.K.; Heinz, E.; Hughes, G.L. CRISPR/Cas9-mediated gene deletion of the ompA gene in symbiotic Cedecea neteri impairs biofilm formation and reduces gut colonization of Aedes aegypti mosquitoes. PLoS Negl. Trop. Dis. 2019, 13, e0007883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Parastan, R.; Kargar, M.; Solhjoo, K.; Kafilzadeh, F. Staphylococcus aureus biofilms: Structures, antibiotic resistance, inhibition, and vaccines. Gene Rep. 2020, 20, 100739. [Google Scholar] [CrossRef]
  136. O’Rourke, J.P.; Daly, S.M.; Triplett, K.D.; Peabody, D.; Chackerian, B.; Hall, P.R. Development of a mimotope vaccine targeting the Staphylococcus aureus quorum sensing pathway. PLoS ONE 2014, 9, e111198. [Google Scholar] [CrossRef] [Green Version]
  137. Novotny, L.A.; Jurcisek, J.A.; Ward, M.O.; Jordan, Z.B.; Goodman, S.D.; Bakaletz, L.O. Antibodies against the majority subunit of type IV pili disperse nontypeable Haemophilus influenzae biofilms in a LuxS-dependent manner and confer therapeutic resolution of experimental otitis media. Mol. Microbiol. 2015, 96, 276–292. [Google Scholar] [CrossRef] [Green Version]
  138. Hwang, G.; Paula, A.J.; Hunter, E.E.; Liu, Y.; Babeer, A.; Karabucak, B.; Stebe, K.; Kumar, V.; Steager, E.; Koo, H. Catalytic antimicrobial robots for biofilm eradication. Sci. Robot. 2019, 4, eaaw2388. [Google Scholar] [CrossRef]
  139. Schairer, D.O.; Martinez, L.R.; Blecher, K.; Chouake, J.S.; Nacharaju, P.; Gialanella, P.; Friedman, J.M.; Nosanchuk, J.D.; Friedman, A.J. Nitric oxide nanoparticles. Virulence 2012, 3, 62–67. [Google Scholar] [CrossRef] [Green Version]
  140. Fulaz, S.; Vitale, S.; Quinn, L.; Casey, E. Nanoparticle–Biofilm interactions: The role of the EPS Matrix. Trends Microbiol. 2019, 27, 915–926. [Google Scholar] [CrossRef]
  141. Kulshrestha, S.; Khan, S.; Hasan, S.; Khan, M.E.; Misba, L.; Khan, A.U. Calcium fluoride nanoparticles induced suppression of Streptococcus mutans biofilm: An in vitro and in vivo approach. Appl. Microbiol. Biotechnol. 2016, 100, 1901–1914. [Google Scholar] [CrossRef]
  142. Jiang, Y.; Geng, M.; Bai, L. Targeting biofilms therapy: Current research strategies and development hurdles. Microorganisms 2020, 8, 1222. [Google Scholar] [CrossRef] [PubMed]
  143. Rocco, C.J.; Davey, M.E.; Bakaletz, L.O.; Goodman, S.D. Natural antigenic differences in the functionally equivalent extracellular DNABII proteins of bacterial biofilms provide a means for targeted biofilm therapeutics. Mol. Oral Microbiol. 2016, 32, 118–130. [Google Scholar] [CrossRef] [PubMed]
  144. Devaraj, A.; Justice, S.S.; Bakaletz, L.O.; Goodman, S.D. DNABII proteins play a central role in UPEC biofilm structure. Mol. Microbiol. 2015, 96, 1119–1135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Novotny, L.A.; Jurcisek, J.A.; Goodman, S.D.; Bakaletz, L.O. Monoclonal antibodies against DNA-binding tips of DNABII proteins disrupt biofilms in vitro and induce bacterial clearance in vivo. EBioMedicine 2016, 10, 33–44. [Google Scholar] [CrossRef] [Green Version]
  146. Freire, M.O.; Devaraj, A.; Young, A.; Navarro, J.B.; Downey, J.S.; Chen, C.; Bakaletz, L.O.; Zadeh, H.H.; Goodman, S.D. A bacterial-biofilm-induced oral osteolytic infection can be successfully treated by immuno-targeting an extracellular nucleoid-associated protein. Mol. Oral Microbiol. 2017, 32, 74–88. [Google Scholar] [CrossRef] [Green Version]
  147. Estellés, A.; Woischnig, A.K.; Liu, K.; Stephenson, R.; Lomongsod, E.; Nguyen, D.; Zhang, J.; Heidecker, M.; Yang, Y.; Simon, R.J.; et al. A high-affinity native human antibody disrupts biofilm from Staphylococcus aureus bacteria and potentiates antibiotic efficacy in a mouse implant infection model. Antimicrob. Agents Chemother. 2016, 60, 2292–2301. [Google Scholar] [CrossRef] [Green Version]
  148. Huang, D.; Wang, J.; Ren, K.; Ji, J. Functionalized biomaterials to combat biofilms. Biomater. Sci. 2020, 8, 4126–4140. [Google Scholar] [CrossRef]
  149. Pelgrift, R.Y.; Friedman, A.J. Nanotechnology as a therapeutic tool to combat microbial resistance. Adv. Drug Deliv. Rev. 2013, 65, 1803–1815. [Google Scholar] [CrossRef]
  150. Ikuma, K.; Decho, A.W.; Lau, B.L.T. When nanoparticles meet biofilms—Interactions guiding the environmental fate and accumulation of nanoparticles. Front. Microbiol. 2015, 6, 591. [Google Scholar] [CrossRef]
  151. Baek, Y.W.; An, Y.J. Microbial toxicity of metal oxide nanoparticles (CuO, NiO, ZnO, and Sb2O3) to Escherichia coli, Bacillus subtilis, and Streptococcus aureus. Sci. Total Environ. 2011, 409, 1603–1608. [Google Scholar] [CrossRef]
  152. Hernández-Sierra, J.F.; Ruiz, F.; Cruz Pena, D.C.; Martínez-Gutiérrez, F.; Martínez, A.E.; de Jesús Pozos Guillén, A.; Tapia-Pérez, H.; Castañón, G.M. The antimicrobial sensitivity of Streptococcus mutans to nanoparticles of silver, zinc oxide, and gold. Nanomed. Nanotechnol. Biol. Med. 2008, 4, 237–240. [Google Scholar] [CrossRef] [PubMed]
  153. Iannitelli, A.; Grande, R.; di Stefano, A.; di Giulio, M.; Sozio, P.; Bessa, L.J.; Laserra, S.; Paolini, C.; Protasi, F.; Cellini, L. Potential antibacterial activity of carvacrol-loaded poly(DL-lactide-co-glycolide) (PLGA) nanoparticles against microbial biofilm. Int. J. Mol. Sci. 2011, 12, 5039–5051. [Google Scholar] [CrossRef] [PubMed]
  154. Kulshrestha, S.; Khan, S.; Meena, R.; Singh, B.R.; Khan, A.U. A graphene/zinc oxide nanocomposite film protects dental implant surfaces against cariogenic Streptococcus mutans. Biofouling 2014, 30, 1281–1294. [Google Scholar] [CrossRef] [PubMed]
  155. Goto, T.; Nakame, Y.; Nishida, M.; Ohi, Y. In vitro bactericidal activities of betalactamases, amikacin, and fluoroquinolones against Pseudomonas aeruginosa biofilm in artificial urine. Urology 1999, 53, 1058–1062. [Google Scholar] [CrossRef]
  156. Carpenter, B.L.; Situ, X.; Scholle, F.; Bartelmess, J.; Weare, W.W.; Ghiladi, R.A. Antiviral, antifungal and antibacterial activities of a BODIPY-based photosensitizer. Molecules 2015, 20, 10604–10621. [Google Scholar] [CrossRef] [Green Version]
  157. Hu, X.; Huang, Y.Y.; Wang, Y.; Wang, X.; Hamblin, M.R. Antimicrobial photodynamic therapy to control clinically relevant biofilm infections. Front. Microbiol. 2018, 9, 1299. [Google Scholar] [CrossRef] [Green Version]
  158. Plaetzer, K.; Krammer, B.; Berlanda, J.; Berr, F.; Kiesslich, T. Photophysics and photochemistry of photodynamic therapy: Fundamental aspects. Lasers Med. Sci. 2009, 24, 259–268. [Google Scholar] [CrossRef]
  159. Li, M.; Li, L.; Su, K.; Liu, X.; Zhang, T.; Liang, Y.; Jing, D.; Yang, X.; Zheng, D.; Cui, Z.; et al. Highly effective and noninvasive near-infrared eradication of a Staphylococcus aureus biofilm on implants by a photoresponsive coating within 20 Min. Adv. Sci. 2019, 6, 1900599. [Google Scholar] [CrossRef] [Green Version]
  160. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; Rodriguez-Torres, M.D.P.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnology 2018, 16, 71. [Google Scholar] [CrossRef] [Green Version]
  161. Motta, J.P.; Wallace, J.L.; Buret, A.G.; Deraison, C.; Vergnolle, N. Gastrointestinal biofilms in health and disease. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 314–334. [Google Scholar] [CrossRef]
  162. Stewart, P.S. Prospects for anti-biofilm pharmaceuticals. Pharmaceuticals 2015, 8, 504–511. [Google Scholar] [CrossRef] [Green Version]
  163. Fleming, D.; Rumbaugh, K. Approaches to dispersing medical biofilms. Microorganisms 2017, 5, 15. [Google Scholar] [CrossRef] [PubMed]
  164. Vuotto, C.; Donelli, G. Novel treatment strategies for biofilm-based infections. Drugs 2019, 79, 1635–1655. [Google Scholar] [CrossRef] [PubMed]
  165. Johnson, C.H.; Spilker, M.E.; Goetz, L.; Peterson, S.N.; Siuzdak, G. Metabolite and microbiome interplay in cancer immunotherapy. Cancer Res. 2016, 76, 6146–6152. [Google Scholar] [CrossRef] [Green Version]
  166. Li, S.; Konstantinov, S.R.; Smits, R.; Peppelenbosch, M.P. Bacterial biofilms in colorectal cancer initiation and progression. Trends Mol. Med. 2017, 23, 18–30. [Google Scholar] [CrossRef] [PubMed]
  167. Le, P.; Kunold, E.; Macsics, R.; Rox, K.; Jennings, M.C.; Ugur, I.; Reinecke, M.; Chaves-Moreno, D.; Hackl, M.W.; Fetzer, C.; et al. Repurposing human kinase inhibitors to create an antibiotic active against drug-resistant Staphylococcus aureus, persisters and biofilms. Nat. Chem. 2020, 12, 145–158. [Google Scholar] [CrossRef]
  168. Li, J.; Zhao, H.; Zheng, L.; An, W. Advances in synthetic biology and biosafety governance. Front. Bioeng. Biotechnol. 2021, 9, 173. [Google Scholar] [CrossRef]
  169. Hong, S.H.; Lee, J.; Wood, T.K. Engineering global regulator Hha of Escherichia coli to control biofilm dispersal. Microb. Biotechnol. 2010, 3, 717–728. [Google Scholar] [CrossRef]
  170. Volke, D.C.; Nikel, P.I. Getting bacteria in shape: Synthetic Morphology approaches for the design of efficient microbial cell factories. Adv. Biosyst. 2018, 2, 1800111. [Google Scholar] [CrossRef]
  171. Fang, K.; Park, O.J.; Hong, S.H. Controlling biofilms using synthetic biology approaches. Biotechnol. Adv. 2020, 40, 107518. [Google Scholar] [CrossRef]
  172. Wood, T.K.; Hong, S.H.; Ma, Q. Engineering biofilm formation and dispersal. Trends Biotechnol. 2011, 29, 87–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Hong, S.H.; Hegde, M.; Kim, J.; Wang, X.; Jayaraman, A.; Wood, T.K. Synthetic quorum-sensing circuit to control consortial biofilm formation and dispersal in a microfluidic device. Nat. Commun. 2012, 3, 613. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biofilm development and antibiofilm strategies. The microbial biofilm cycle could be classified into 4 phases: Initial attachment, Adhesion, Maturation, and Dispersal. The biofilm inhibitory and dispersal strategies are summarized as per the stages in biofilm development. (A) The initial attachment can be disrupted by interfering with the interactions between the surface and the microorganism either by surface remodeling or physical removal of the biofilms; (B) Adhesion can be inhibited by targeting biofilm EPS and cellular division; (C) Disruption of biofilms in proliferating and maturing phase may be accomplished either by physical removal or by damaging the EPS matrix primarily by affecting the formation of pathogenic microenvironments (such as hypoxia or low pH), and quorum sensing along with the eradication of persister cells. (D) Biofilm dispersal could be achieved by remodeling the EPS matrix or accelerating the dispersal mechanisms. (Different colors of the cells represent different bacteria within the biofilm. Circular cells represent cocci and rod shaped cells represent bacilli).
Figure 1. Biofilm development and antibiofilm strategies. The microbial biofilm cycle could be classified into 4 phases: Initial attachment, Adhesion, Maturation, and Dispersal. The biofilm inhibitory and dispersal strategies are summarized as per the stages in biofilm development. (A) The initial attachment can be disrupted by interfering with the interactions between the surface and the microorganism either by surface remodeling or physical removal of the biofilms; (B) Adhesion can be inhibited by targeting biofilm EPS and cellular division; (C) Disruption of biofilms in proliferating and maturing phase may be accomplished either by physical removal or by damaging the EPS matrix primarily by affecting the formation of pathogenic microenvironments (such as hypoxia or low pH), and quorum sensing along with the eradication of persister cells. (D) Biofilm dispersal could be achieved by remodeling the EPS matrix or accelerating the dispersal mechanisms. (Different colors of the cells represent different bacteria within the biofilm. Circular cells represent cocci and rod shaped cells represent bacilli).
Pathogens 11 00292 g001
Figure 2. Diagrammatic representation of antibiotic (drug) tolerance in biofilms. Possible tolerance mechanisms at (A) community and (B) cellular level.
Figure 2. Diagrammatic representation of antibiotic (drug) tolerance in biofilms. Possible tolerance mechanisms at (A) community and (B) cellular level.
Pathogens 11 00292 g002
Figure 3. Recent approaches for biofilm treatment. Biofilms could be controlled by using (A) Antibiofilm agents that target various compounds involved in biofilm formation (B) Therapeutic methods directly targeting the biofilm formation process.
Figure 3. Recent approaches for biofilm treatment. Biofilms could be controlled by using (A) Antibiofilm agents that target various compounds involved in biofilm formation (B) Therapeutic methods directly targeting the biofilm formation process.
Pathogens 11 00292 g003
Table 1. Classes of antibiofilm agents and their functions.
Table 1. Classes of antibiofilm agents and their functions.
Antibiofilm AgentsFunctions
Class Ipenetrate the biofilm EPS and decrease the growth of cells
Class IIinterfere with the adherence of bacteria and the formation of biofilm phenotype
Class IIIcontrols both the growth of bacteria with biofilm phenotype as well as the EPS synthesis
Class IVdisperse the mature biofilms
Table 2. Different types of antimicrobial agents and their mode of action.
Table 2. Different types of antimicrobial agents and their mode of action.
Name of the CompoundTypeMode of ActionEffective AgainstReference
Triton X-100surfactantautolysis, targeting EPSS. aureus[31,32]
Tween 80surfactantN/AP. aeruginosa, S. aureus[30]
Quarternary ammonium compoundssurfactantCell lysis and deathseveral bacteria[34]
Poloxamer containing non-ionic surfactantsurfactantEPS metalloproteinase modulationP. aeruginosa[34]
Rhamnolipidsbio-surfactantN/AS. aureus, Salmonella Enteritidis, and Listeria monocytogenes[36]
EDTAchelatorsdamage to cell wallP. aeruginosa[40]
Chitosanbiomaterialmembrane damageP. aeruginosa[41]
Secondary metabolite from Citrus limonoidssecondary metabolitequorum sensingVibriyo harveyi[42]
Cyclo(l-Tyr-l-Leu)secondary metaboliteinhibit EPSS. epidermidis[43]
Cahuitamycinssecondary metaboliteN/AA. baumanii[44]
Phlorotanninsecondary metabolitedamaging membrane permeability/ cell lysisMRSA[45]
α-amylaseenzymedegrade EPSMRSA[46]
Polyamine norspermidinepolyamineinteracts with EPSB. subtilis, E. coli and S. aureus[47]
D-amino acidsamino acidtarget YqxME. coli, S. aureus[48]
N-acetylcysteine/NACamino aciddegrade EPS polysaccharideRapidly growing Mycobacterium[49]
Esp (Serine protease)enzymesdegrade EPS protein contentS. aureus[50]
DNase Ienzymesdegrade eDNAE. coli, S. aureus[51]
tea-tree oilsecondary metabolitemetabolismS. aureus[52]
Protease from P. aeruginosaenzymesdegrade EPS protein contentS. aureus[53]
Table 3. Natural compounds as antiquorum sensing molecules in biofilm dispersal.
Table 3. Natural compounds as antiquorum sensing molecules in biofilm dispersal.
Compound/MoleculeMode of ActionEffective AgainstReference
Garlic extractsinhibits QSP. aeruginosa[62]
Garlic extractsinhibit LasR and LuxRP. aeruginosa[63]
Isolimonic acidcell-cell signalingE. coli[64,65]
Isolimonic acidreduce LuxR DNA bindingVibrio spp.[65]
Cinnamaldehydeswimming motilityE. coli[66]
Hordeninedecrease in signaling molecule, inhibition of QS-related genesP. aeruginosa[67,68]
Autoinducing peptide type I (AIP-I)inhibit QSS. aureus[74]
RNAIII-inhibiting peptide (RIP)inhibit QSS. aureus[75]
Querentindecrease LasI/R, RhlI/R expressionsP. aeruginosa[73]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shrestha, L.; Fan, H.-M.; Tao, H.-R.; Huang, J.-D. Recent Strategies to Combat Biofilms Using Antimicrobial Agents and Therapeutic Approaches. Pathogens 2022, 11, 292. https://doi.org/10.3390/pathogens11030292

AMA Style

Shrestha L, Fan H-M, Tao H-R, Huang J-D. Recent Strategies to Combat Biofilms Using Antimicrobial Agents and Therapeutic Approaches. Pathogens. 2022; 11(3):292. https://doi.org/10.3390/pathogens11030292

Chicago/Turabian Style

Shrestha, Looniva, Hai-Ming Fan, Hui-Ren Tao, and Jian-Dong Huang. 2022. "Recent Strategies to Combat Biofilms Using Antimicrobial Agents and Therapeutic Approaches" Pathogens 11, no. 3: 292. https://doi.org/10.3390/pathogens11030292

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