Next Article in Journal
Characterizing the Potato Growing Regions in India Using Meteorological Parameters
Next Article in Special Issue
Antibiofilm Activity of LL-37 Peptide and D-Amino Acids Associated with Antibiotics Used in Regenerative Endodontics on an Ex Vivo Multispecies Biofilm Model
Previous Article in Journal
SARS-CoV-2 and Influenza Vaccines in People with Excessive Body Mass—A Narrative Review
Previous Article in Special Issue
Comparative Study of Antibacterial, Antibiofilm, Antiswarming and Antiquorum Sensing Activities of Origanum vulgare Essential Oil and Terpinene-4-ol against Pathogenic Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Strategies as Potential Weapons against Bacterial Biofilms

1
Department of Food Hygiene and Technology, Faculty of Veterinary Medicine, Afyon Kocatepe University, Afyonkarahisar 03200, Turkey
2
Department of Animal Production and Veterinary Public Health, Faculty of Veterinary Medicine, University of Life Sciences “King Michael I” from Timișoara, 300645 Timisoara, Romania
3
Department of Parasitology and Dermatology, University of Life Sciences “King Michael I” from Timișoara, 300645 Timisoara, Romania
4
Department of Animal Nutrition and Nutritional Diseases, Faculty of Veterinary Medicine, Afyon Kocatepe University, Afyonkarahisar 03200, Turkey
5
Department of Chemistry and Chemical Engineering, SBA School of Science & Engineering (SBASSE), Lahore University of Management Sciences (LUMS), Lahore 54792, Pakistan
6
Department of Biochemistry, Faculty of Veterinary Medicine, Afyon Kocatepe University, Afyonkarahisar 03200, Turkey
7
Department of Food Hygiene and Technology, Faculty of Veterinary Medicine, Kyrgyz-Turkish Manas University, Bishkek KG-720038, Kyrgyzstan
8
Centre de Recherche en Technologies Agro-Alimentaires, Campus Universitaire Tergua Ouzemmour, Bejaia 06000, Algeria
9
Laboratory of Entomology and Agriculture Zoology, Department of Agriculture, Crop Production and Rural Environment, University of Thessaly, 38446 Volos, Greece
10
Department of Animal Husbandry, Faculty of Veterinary Medicine, Trakia University, Students’ Campus, 6015 Stara Zagora, Bulgaria
11
Department of Internal Medicine, Faculty of Veterinary Medicine, University of Life Sciences “King Michael I” from Timișoara, 300645 Timisoara, Romania
12
Department of Infectious Disease and Preventive Medicine, Faculty of Veterinary Medicine, University of Life Sciences “King Michael I” from Timișoara, 300645 Timisoara, Romania
13
National Center for Veterinary Drug Safety Evaluation, College of Veterinary Medicine, China Agricultural University, Beijing 100193, China
*
Author to whom correspondence should be addressed.
Life 2022, 12(10), 1618; https://doi.org/10.3390/life12101618
Submission received: 29 September 2022 / Revised: 7 October 2022 / Accepted: 12 October 2022 / Published: 17 October 2022
(This article belongs to the Special Issue Antibiotic Resistance in Biofilm)

Abstract

:
Microbial biofilm is an aggregation of microbial species that are either attached to surfaces or organized into an extracellular matrix. Microbes in the form of biofilms are highly resistant to several antimicrobials compared to planktonic microbial cells. Their resistance developing ability is one of the major root causes of antibiotic resistance in health sectors. Therefore, effective antibiofilm compounds are required to treat biofilm-associated health issues. The awareness of biofilm properties, formation, and resistance mechanisms facilitate researchers to design and develop combating strategies. This review highlights biofilm formation, composition, major stability parameters, resistance mechanisms, pathogenicity, combating strategies, and effective biofilm-controlling compounds. The naturally derived products, particularly plants, have demonstrated significant medicinal properties, producing them a practical approach for controlling biofilm-producing microbes. Despite providing effective antibiofilm activities, the plant-derived antimicrobial compounds may face the limitations of less bioavailability and low concentration of bioactive molecules. The microbes-derived and the phytonanotechnology-based antibiofilm compounds are emerging as an effective approach to inhibit and eliminate the biofilm-producing microbes.

1. Introduction

A biofilm is a complex of micro-organisms that sustains a structured and organized pathway for their growth, proliferation, and survival on any surface [1,2]. A single bacterial species may develop the biofilm organizations, or the biofilm can also be of mixed microbial species adhered to a surface [3]. The survival of biofilm-forming bacterial cells depends upon the alignment of extracellular polymeric substance-encapsulated micro-colonies in the matrix [4]. The biofilm-forming microbial cells maintain their micro-environment by controlling temperature, nutrients, and pH, which can affect biofilm formation. Biofilms are known to cause several infections or diseases in humans [5,6]. Several antibiotics have been used against biofilm-associated infections, but increased antimicrobial resistance (AMR) has directly been linked to biofilm microbes [7]. This inefficacy of numerous antibiotics against several biofilm-linked infections has gradually enhanced the emergence of AMR.
The development of microbial resistance to antibiotics reduces or inhibits the efficacy of antibiotics. Consequently, it has been demonstrated that the improper use of antibiotics leads to the emergence of antibiotic resistance [8]. Many antibiotics are losing efficacy due to several micro-organisms’ expeditious developments in multidrug resistance. Several factors responsible for causing AMR and intrinsic biofilm formation have been recognized as major critical factors [9]. Moreover, AMR caused by biofilm development may cause harmful recurrent chronic microbial infections. Therefore, the discovery of a significant treatment approach is much needed to combat AMR. Several researchers are designing and evaluating new combat strategies based on one health concept [10].
This study explores the new combat approaches that avoid existing resistance mechanisms. Using different natural products, such as plant-derived components, bee products, marine-derived components, and plant-based nanomaterials, is gaining great attention from researchers to design novel antibiofilm compounds to avoid AMR. In the current review, the emerging global concerns regarding the development of biofilm resistance and their control by describing some potential therapeutic compounds have been recapitulated.

2. Biofilm Formation

Biofilm can be defined as the complex aggregation of micro-organisms, firmly adhered to a surface and micro-colonized into extracellular polymeric substances (EPS) matrix. EPS is comprised of exopolysaccharides, nucleic acids, lipids, and proteins [11]. The emergence of biofilm formation is a multistep process in which EPS performs particularly required functional and structural roles. Microbial communities may attach to both abiotic and biotic surfaces facilitated by EPS. In order to maintain a biofilm lifestyle, the EPS matrix provides essentially required chemical microenvironments and mechanical stability [12].
In addition, the EPS also improves biofilm tolerance toward different antimicrobial agents and immune cells. The biofilm maturation needs several developmental phases with particular features, resulting in biofilm development on the surfaces. Additionally, understanding every developmental phase is essential for designing and applying proper antimicrobial agents against microbial biofilms. Each developmental phase is briefly elaborated on in Figure 1.

2.1. Surface Adhesion

Biofilm development is initiated with the attachment of microbial cells to any surface. Several sensing pathways (e.g., BasSR, BaeSR, and CpxAR) facilitate the microbes’ identification of the surface [13]. After initial adhesion, the microbial cells start to multiply and divide expeditiously under favorable conditions [14]. The microbial cells produce adhesins (enzymes) that facilitate their attachment to the host surfaces. This phase explicitly targets the adherence mechanism between the microbes and the surfaces. Thus, in this phase, significant combating intervention can be achieved by entirely disrupting the attachment mechanisms of micro-organisms with the surfaces via mainly targeting cell surface-linked adhesins. Consequently, the initial biofilm development can be inhibited by disrupting the initial adhesion process [15].
Surface adhesion is essentially the required parameter for biofilm development. The adhesion initiation and the biofilm dispersal start with the adherence capacity of a specific bacterial species concerning the host surface. The surface attachment, along with biofilm development, is the survival strategy of the microbes that typically adhere to themselves in a particular way to meet their environmental and nutritional requirements. The surface adhesion consists of two phases: the primary/reversible phase and the secondary/irreversible attachment phase [16]. Both of these phases are entirely controlled by the gene’s expression. The attachment processes require several points to be considered, including microbial species, environmental conditions, gene products, and surface composition [17]. In the reversible phase, microbes hydrophobically interact with the abiotic surfaces for adhesion, whereas adherence with biotic surfaces takes place by developing molecular interactions [18].

2.2. Biofilm Maturation

The formation of mature biofilm is followed by early biofilm development when the microbes begin to multiply and divide, creating micro-colonies incorporated with the EPS matrix [19]. The EPS matrix plays a multifunctional role, allowing the establishment of several physical and chemical microhabitats facilitating the microbes to build social and polymicrobial interactions. The established biofilms can be disrupted or removed by employing several targeting approaches, including the EPS matrix disruption, targeting microenvironments (such as hypoxia or low pH), physical removal, targeting polymicrobial interactions, and eliminating dormant cells, etc., providing a great scope for designing biofilm combating antimicrobial therapeutics [20].

2.3. Biofilm Dispersal

During the dispersal phase, the micro-organisms of sessile biofilm begin to disperse and change into motile form. On the other hand, the microbes that do not produce extracellular polysaccharides directly scatter into the environment by applying a mechanical force. The dispersed microbial communities produce saccharolytic enzymes, which release surface microbes towards a new place for colonization. The Escherichia coli produces N-acetyl-heparosan lyase and Pseudomonas aeruginosa, P. fluorescens produces alginate lyase, and Streptococcus equisimilis produces hyaluronidase. The development of flagella originates with the upregulation of protein expression by microbial cells. It allows the microbes to move to a new place for colonization and supports them in spreading biofilm-associated diseases. The remodeling of the EPS matrix and the dispersal pathways activation may induce dispersion of biofilm that can assist in combating biofilms [18].

3. Biofilm Composition

Biofilm is a complex of variability and heterogeneity consisting of 10–25% microbial cells and a 75–90% self-developed EPS matrix [21]. Moreover, the water channels or interstitial voids of microbial biofilms are mandatory for micro-colonies’ separation from each other [22]. The EPS creates a covering scaffold that grips the biofilm cells together and facilitates communication between cell-to-cell, providing cohesive and adhesive forces for biofilm development. EPS facilitates nutrient availability, maintaining the deoxyribonucleic acid (DNA) availability for horizontal gene transfer and provides a defensive barrier against antibiotics, desiccation, oxidizing biocides, host defense immune system, and ultraviolet radiation [23]. The EPS constituents include polysaccharides, extracellular proteins, extracellular DNA, lipids, surfactants, and water.

3.1. Polysaccharides

Polysaccharides developed by microbes can be categorized into two categories: hetero-polysaccharides and homo-polysaccharides. Most polysaccharides are heterogeneous in nature, while only a few are homogenous in nature, such as glucans, sucrose-based fructans, and cellulose [24]. Several interactions, such as electrostatic interactions, van der Waals forces, ionic interactions, and hydrogen bonding, help in polysaccharide interactions with themselves or ions and proteins to maintain the architecture of biofilms [25]. The primary function of polysaccharides is to provide a protective role by mediating microbial adhesion among the micro-colonies and maintaining the structural stability of biofilms [26]. Three types of exopolysaccharides, such as alginate, Pel, and Psl, contribute to biofilm development and maintain the structural stability of P. aeruginosa biofilms [27]. Some polysaccharides identified in different microbial biofilms are represented in Table 1.

3.2. Extracellular Proteins

The biofilm complex can have a substantial number of extracellular proteins [34]. Their interaction with exopolysaccharides and nucleic acids facilitates surface colonization and stabilization in the biofilm matrix [35]. Some proteins mediate the degradation and dispersion of biofilm matrix, such as glycosyl hydrolase, dispersin B induces polysaccharide degradation [36], proteases dissolve the proteins of matrix [37], and some DNases cause extracellular nucleic acid breakage [38]. Toyofuku et al. described that 30% of EPS-matrix proteins in P. aeruginosa were observed in outer-membrane vesicles as membrane proteins. Some of them were secreted and lysed by cell-derived proteins [39].
Several extracellular enzymes have also been found in microbial biofilms; some of them are involved in biopolymer degradation. The extracellular enzyme substrates contain water-insoluble components (such as lipids, cellulose, and chitin), water-soluble compounds (such as proteins, polysaccharides, and nucleic acids), and the biofilm entraps organic particles [40]. Additionally, some extracellular enzymes can be used for EPS structural degradation to mediate the microbial detachment in biofilms.

3.3. Extracellular DNA (eDNA)

eDNA is one of the prime components of the EPS matrix, which is essential for the accumulation of microbes within the biofilm. The amount of eDNA production may vary even among closely linked microbial species. eDNA is the primary structural constituent in the matrix of a Staphylococcus aureus biofilm, while in S. epidermidis biofilms, it is produced as a minor constituent [41]. It has been revealed that eDNA is a vital component of the biofilm matrix and its mode of life [42,43]. eDNA has also been found as a major component in the biofilm matrix of P. aeruginosa and facilitates intercellular interactions [44].
Moreover, the eDNA was observed to inhibit the P. aeruginosa biofilm formation, while in Bacillus cereus the eDNA acts similar to adhesins [45,46]. Okshevsky and Meyer reported that eDNA is involved in cell adhesion, structural stability maintenance, and horizontal gene transfer and also protects the immune system and antimicrobials [47]. eDNA has been observed to facilitate cell adhesion and biofilm development in Listeria monocytogenes [48]. Wilton et al. found that eDNA causes acidification of the biofilm matrix and consequently enhances resistance of P. aeruginosa biofilms against different antibiotics [49].

3.4. Surfactants and Lipids

The extracellular polysaccharides, proteins, and eDNA are quite hydrated (hydrophilic) molecules, while the other components of EPS exhibit hydrophobic properties. Some bacterial species, such as Rhodococcus spp. generate hydrophobic EPS, and can attach to Teflon and may colonize the waxen surfaces with the help of hydrophobic EPS [50]. In the EPS matrix, a few lipids exhibiting surface active properties, such as surfactin, viscosin, and emulsan, increase the availability of hydrophobic molecules by causing their dispersion [24]. An important surfactant class, biosurfactants, begins the micro-colony formation, aids in biofilm structural integrity, and mediates biofilm dispersal [51].

3.5. Water

Water is recognized as the largest constituent (accounts for up to 97%) of the EPS matrix of most of the biofilms, and it retains the biofilm hydrated and protects against desiccation [24]. The water in the biofilm matrix may exist in the form of solvents or can also be bound inside the bacterial cells’ capsules [52]. The binding and movement of water inside the biofilm matrix are essential to diffusion mechanisms that occur inside the biofilm and result in fine biofilm structure development [53]. The amount of available water is responsible for nutrient flow and availability within the microbial biofilms [52].

4. Biofilm Structural Stability Parameters

The antimicrobial resistance and the other functional characteristics of microbial biofilms are linked with the structure of the biofilm, matrix shape, and the 3D organization of microbes [54]. The local environmental heterogeneous conditions inside the biofilm matrix impact the microbial gene expression and the metabolic actions of biofilm-developing microbial cells [55,56]. The closely packed microbial cells and the water channels are the two major constituents of biofilm formation [57]. The structural familiarity of microbial biofilms is of greater concern in identifying their behavioral and survival strategies. The biofilm architecture and formation variability have been analyzed by applying specific parameters, such as substratum exposure, bio-volume, thickness, and roughness, and observing significant inter- and intra-species variability [56].
The cell-cell communication, environmental influences, and the secondary messengers, such as c-di-GMP and cAMP, structure the biofilms by providing microbes with better environmental adaptability [58]. Several other factors influence the biofilm architecture, such as nutrient availability, microbial motility, hydrodynamic conditions, exopolysaccharides and protein abundance, and anionic and cationic concentrations within the biofilm. In P. aeruginosa, an EPS known as alginate facilitates biofilm formation and its architectural stability [59].
The EPS in Vibrio cholera and E. coli facilitates biofilm development in a three-dimensional configuration [60,61]. An EPS and a secreted protein called TasA are vital for developing a fruiting body, such as a Bacillus subtilis biofilm, and maintaining the biofilm matrix’s integrity [62]. The architecture of biofilm can be altered by the exopolysaccharides’ substituents, such as acetyl groups, known to be responsible for enhanced cohesive and adhesive biofilm properties [24].

4.1. Proteins

Manifoili et al. determined the impact of three different mitogen-activated protein kinases (MAPKs) (such as SakA, MpkA, and MpkC) and protein phosphatases (PhpA) on Aspergillus fumigatus biofilm formation. MAPKs reduce the A. fumigatus adhesion in biofilm formation. The ΔpphA strain was found to be more susceptible to cell wall destroying antimicrobials, had less chitin, and enhanced β-(1,3)-glucan, resulting in reduced adherence and biofilm formation [63].
The bacterial cell wall-linked fibronectin-binding proteins (FnBPs) (FnBPB and FnBPA) mediate the biofilm development of methicillin-resistant S. aureus (strain LAC) by promoting bacterial adhesion and accumulation [64]. The outer membrane protein W (OmpW) contributes to Cronobacter sakazakii survival and biofilm formation under NaCl-stressed conditions [65]. A surface protein, BapA1, plays a significant role in bacterial adhesin and biofilm formation. BapA1 carries the nine putative pilin iso-peptide linker domains, which are significant for bacterial accumulation of pilus in several Gram-positive bacteria, such as Streptococcus parasanguinis [66].

4.2. Genes and Signaling Cascades

Intracellular cyclic dinucleotide and extracellular quorum sensing (QS) signaling cascades are crucial in biofilm development. It has been reported that these two signaling pathways may coincide or link up and synergistically mediate biofilm formation [67]. QS is the process of intercellular communication that enables the bacteria to adapt to harsh environmental conditions. They mediate biofilm formation by activating small signaling molecules, such as autoinducer-2 (AI-2), auto-inducing peptide (AIP), and N-acyl-homoserine lactones (AHL), in Gram-positive and -negative bacteria, respectively [68]. AI-2 mediates the QS and biofilm development with bhp- and ica-dependent modes. AI-2 regulates the QS in S. epidermidis through increased transcription levels of bhp (biofilm-linked protein containing icaR) and ica operon [69].
QS signaling cascades observed in P. aeruginosa include integrated QS (IQS), PQS, rhl, and las. These QS systems activate each other by regulating QS-associated genes [70]. The small non-coding RNAs (sRNAs) regulate the bacterial transition from planktonic-sessile bacterial biofilms. The two-component regulatory systems (TCSs), such as RsmZ and RsmY targeting RsmA, are involved in P. aeruginosa biofilm formation [71]. Several sRNAs have been studied that regulate the activity or expression of different transcriptional regulators to mediate bacterial adhesion and increase biofilm formation (Table 2).

5. Resistance Mechanism in Biofilms

Several antimicrobial resistance mechanisms have been identified, and biofilm development is one of the major factors of resistance emergence. Mechanisms allowing microbial biofilms to resist or tolerate the antimicrobials’ actions are discussed here.

5.1. The Structural Complexity of Biofilms

As EPS is crucial for a biofilm’s architectural stability, it also acts as a physical barrier, protecting or shielding the embedded microbes against antimicrobials, ultraviolet light, etc. [85,86]. The polysaccharides (negatively charged) can significantly bind to the aminoglycoside antibiotics (positively charged) and block their penetration [87]. The EPS barrier can reduce the diffusion of small compounds, such as H2O2 (hydrogen peroxide), to microbial cells inside the biofilm. It has been observed that P. aeruginosa in the planktonic state is more susceptible to H2O2, while in the form of a biofilm, it can survive even at a very high concentration of H2O2 [88].

5.2. The Heterogeneity of Biofilms

The heterogeneity inside the developed biofilms averts the entire eradication of all involved microbial cells by antimicrobials. There are oxygen and nutrient gradients from the top-bottom of microbial biofilms. From the top-bottom of the biofilm matrix, the oxygen and nutrient reduction leads to a reduced growth rate and metabolic activity [89]. The protein expression of bacterial cells in biofilm is diverse and quite different from planktonic cells, which may also contribute to microbial resistance development. For example, in its biofilm form, a rice endophytic bacterium, Pantoea agglomerans YS19, expresses SPM43.1 protein (acid-resistant) at high levels to resist harsh environmental conditions [90,91].

5.3. Quorum Sensing

Quorum sensing (QS) in microbes is a cell-cell communication mediated by activating specific signaling molecules, facilitating environmental adaptation to microbes [2]. QS is a crucial mechanism for regulating and developing biofilms by reducing or inhibiting the effectiveness of antimicrobials against biofilm bacteria [92]. Gram-negative and positive bacterial species communicate using these signaling molecules, also known as autoinducers (AIs) [2]. Some QS signaling molecules used by Gram-positive and negative bacteria include (a) N-acyl homoserine lactone (AHL), (b) autoinducer-2 (AI-2) by V. harveyi, (c) autoinducing peptide 1 (AIP-1) by S. aureus, (d) N-(3-oxoacyl)-l-homoserine lactone (3-oxo-AHL), (e) diffusible signaling factor (DSF), (f) N-(3-hydroxyacyl) homoserine lactone (3-hydroxy-AHL), (g) 2-heptyl-3-hydroxy-4(1H)-quinolone by P. aeruginosa, and (h) hydroxy-palmitic acid methyl ester (PAME). They are presented in Figure 2.
Environmental factors, such as nutrient deficiency, pH, antimicrobials, and salt concentrations, regulate QS-mediated activity in bacterial biofilms [93]. The feed-forward mechanism enhances the QS communication in the biofilm matrix [94]. The QS signals facilitate biofilm formation when their concentration reaches the threshold [95]. Subsequently, the QS signaling molecules are translated into cells for gene expression modulation. These genes are crucial for environmental adaptation, leading to biofilm formation [96]. The QS signaling system regulates bacterial and EPS secretion systems [97] and multidrug efflux pumps [98]. The complete information about QS signaling molecules may offer significant information for developing novel methods or chemicals for combating microbial biofilms. This research era has grabbed the great attention of researchers for designing and developing significant molecules with the ability to neutralize or compete with the QS signaling molecules or their receptors [99].

5.4. Enhanced Efflux Pumps

The efflux pumps (proteinaceous) entrenched in the cytoplasmic membranes act as active transporters. The multidrug and toxic compound extrusion family (MATE), the ATP-binding cassette family (ABC), the small multidrug resistance family (SMR), the resistance-nodulation-division family (RND), and the major facilitator superfamily (MF) are commonly reported efflux pump classes in different bacteria [100,101].
Several molecular studies have revealed that increased efflux pumps are a popular and criticizing resistance mechanism in microbial biofilms [102]. This mechanism has been extensively studied in a commonly found biofilm-producing P. aeruginosa pathogen [103]. The PA1874-1877 (cluster of genes) involved in developing resistance in biofilms was discovered by Zhang and Mah. Overexpression of PA1874-1877 in biofilm cells facilitates resistance in a biofilm-specific manner [104]. Numerous efflux pump genes inducing biofilm-specific resistance through their overexpression have been identified. For example, in RND-3 efflux pumps, the overexpression of BCAL1672-1676 induces biofilm resistance against ciprofloxacin and tobramycin, while in RND-8 and RND-9, the overexpression of BCAM0925-0927 and BCAM1945-1947 provides resistance to Burkholderia cepacia against tobramycin [105].

6. Pathogenicity of Biofilm Microbes

Numerous factors are known to contribute to pathogenicity in biofilm-producing microbes. Microbial biofilms release different extracellular substances, altering the gene regulation of several microbial virulence factors. Moreover, the biofilm-producing microbes strengthen the maturation rate of biofilms to escape from host defenses, enhance the activity of β-lactamase, and for plasmid-mediated gene transfer resulting in intense virulence and antimicrobial resistance with enhanced mutation rate and efflux pump. The properties of the extracellular matrix contribute to the biofilm’s pathogenicity, offering a defensive barrier with less antimicrobial and immune cell penetration [106].
The MIC of antibiotics is significant against planktonic microbes but not effective against biofilm microbes [107]. Microbial biofilms can induce several persistent biofilm-associated infections, such as urinary tract infections, middle-ear infections, dental caries, endocarditis, cystic fibrosis, osteomyelitis, and implant-induced infections. Numerous pathogenic microbes are involved in causing persistent biofilm infections; some of them are listed in Table 3.
The pathogenic activity of microbes in the form of biofilms is significantly higher, and they can escape from host defense cells and antibiotics. In numerous infections, biofilm bacteria are concerned with the pathogenesis and clinical symptoms [114]. Opportunistic pathogenic bacteria, such as P. aeruginosa and S. aureus, can cause chronic biofilm infections, and hospitalized individuals (approximately 8–10%) are more vulnerable to carrying infections.

6.1. Health Problems and Infections Caused by Biofilm Bacteria

Biofilm-associated infections pose a threat to human health. Over the last few decades, innovative methods have been discovered to control microbial infections. Biofilm formation in the era of the food industry poses a serious threat to human health. Biofilms may contain only one type of bacteria, different bacterial species, or fungal species that may be pathogenic and may only target immunocompromised patients (cancer patients, organ recipients, HIV patients, etc.). Systematic diseases (E. coli, L. monocytogenes), food intoxication (P. aeruginosa, S. aureus, B. cereus), and gastroenteritis (Salmonella enterica, E. coli) can be caused by biofilm-producing pathogens [114].

6.2. Biofilms in the Food Industry

Foodborne infections may arise from microbial biofilm development on food processing equipment or food matrices. Biofilms formed on food processing equipment can secrete toxins and may result in food poisoning. Biofilm formation in any food industry may put human health at potential risk. The severity of the risk is directly dependent on the microbial species of the biofilm matrix.
Food processing plants provide suitable conditions for biofilm development on food surfaces due to the complexity of manufacturing or processing plants, mass product yield, long manufacturing durations, and large biofilm formation areas [115]. These biofilm formations may contribute to the emergence of biofilm-associated foodborne infections. Approximately 80% of microbial infections in the USA are considered to be specifically associated with biofilm-producing foodborne pathogens [116]. Mixed-species or polymicrobial biofilm formation is a highly diverse phenomenon and depends upon environmental conditions [117], adherence characteristics of surfaces [118], involved microbial cells [119], and components of the food matrix [120]. Adherence surface characteristics, such as electrostatic charge, topography, interface roughness, and hydrophobicity, impact biofilm development and consequently affect the surface’s hygienic status [118,121].
Properties of microbial cells, such as components of cell membranes (e.g., lipopolysaccharides and proteins), hydrophobicity, exopolysaccharides (EPS) production by microbes, and the bacterial appendages (e.g., fimbriae, pili, and flagella), contribute to a crucial role in biofilm formation [118]. Some studies have reported that microbial adherence is more likely to develop on rough surfaces [122], and some experiments indicated no association between microbial adherence and surface roughness [123]. The components of the food matrix in food processing plants may influence microbial adhesion, such as food waste, e.g., carbohydrates, proteins, and fat-enriched meat and milk exudates, mediate microbial growth and proliferation, and facilitate the dual-species biofilm development by S. aureus and E. coli [124,125]. Biofilm-producing foodborne pathogens have emerged as a serious threat to human health. Some foodborne pathogens with biofilm-forming ability and their harmful effects are listed in Table 4. Biofilms have been found to be associated with different outbreaks or epidemics (Table 5).

7. Biofilm Control

The global rise in antibiotic resistance has led to the failure of antibiotics. The ABR has become a major threat to human health. Therefore, alternative therapies have been reported to eliminate or inhibit biofilm formation and their associated infections. Different points can be targeted for biofilm inhibition and eradication at different stages of biofilm formation (Figure 3). These combating strategies include inhibition of planktonic cells, inhibition of bacterial adhesion, surface alteration, biofilm removal, degradation of EPS, QS inhibition, dispersion of biofilms, and matrix degradation.

8. Biofilm Controlling Compounds

Many natural compounds can act as biofilm-controlling compounds by interfering with QS, possessing antiadhesive properties, and inhibiting biofilm formation (Figure 4).

8.1. Natural Plants and Bee Products

Several naturally occurring compounds can be used as antibiofilm molecules to eradicate or inhibit biofilm development. The garlic extract can significantly block QS and may promote rapid virulence attenuation (e.g., elastase, protease A, exo- and cytotoxin production or motility, and adhesion capacity reduction) of P. aeruginosa by polymorphonuclear leukocytes (PMNs) within the immune response of a mouse infection model [146]. Persson et al. synthesized some QSIs derived from garlic extracts and AHLs [147]. Chamaemelum nobile is a naturally occurring, well-known plant for its antimicrobial, anti-inflammatory, antiseptic, spasmolytic, anticatarrhal, sedative, and carminative properties. It can inhibit P. aeruginosa biofilm by disrupting QS [148]. Proanthocyanidins extracted from cranberries have significantly inhibited the adhesion of E. coli to uroepithelial cells [149]. Cranberry juice also significantly inhibited the site-specific adherence of Helicobacter pylori to the gastric mucous of humans [150], and it also prevented Streptococci spp. biofilm formation [151,152]. Eighty medicinal plants, more prominently Fritillaria verticillata, Rhus verniciflua, Cocculus trilobus, and Liriope platyphylla, have been analyzed for their antibiofilm activity. Comparatively, Cocculus trilobus (ethyl acetate fraction) has shown the highest antibiofilm activity against Gram-positive bacteria by providing effective antiadhesive activity [153]. A Chinese herb, Herba patriniae, with medicinal properties, has averted the gene expression of six genes linked with biofilm development and EPS production in P. aeruginosa [154]. The Ginkgollic acid isolated from a plant, Ginkgo biloba, exhibited antitumor, antimicrobial, and neuroprotective and is active against S. aureus strains and E. coli biofilm formation [155,156].
Elekhnawy et al. determined the antiquorum sensing and biofilm inhibitory potentials of Dioon spinulosum plant extract against the clinical isolates of P. aeruginosa. The in vitro analysis of antibiofilm activity showed a 77.1–34.3% reduction in biofilm formation at 250–500 μg/mL concentrations. Both in vitro and in vivo investigations revealed a significant reduction in P. aeruginosa biofilm formation. However, preclinical studies leading to clinical studies are recommended to allow its practical application in treating P. aeruginosa infections [157]. Obaid et al. studied the antibiofilm activity of six plant extracts, such as Apium graveolens, Plantago ovata, Vitis vinifera, Viscus album, Senna acutifolia, and Melissa officinalis, against Aggregatibacter actinomycetemcomitans. The A. actinomycetemcomitans was collected from patients with dental caries. The obtained results indicated that the A. actinomycetemcomitans was more sensitive to S. acutifolia and M. officinalis, with zone inhibitions of 33 and 35 mm, respectively [158]. Negam et al. evaluated the antifungal and antibiofilm potential of Encephalartos laurentianus (methanol extract) against C. albicans. The in vitro antibiofilm analysis revealed a 62.5–25% reduction in C. albicans cell percentage. The in vivo evaluations of E. laurentianus performed on C. albicans infected rats resulted in an increased survival rate with a protective effect against renal damage caused by C. albicans [159].
Olawuwo et al. investigated the in vitro antibiofilm activity of Acalypha wilkesiana, Alchornea laxiflora, Ficus exasperata, Jatropha gossypiifolia, Morinda lucida, and Ocimum gratissimum plant extracts against poultry pathogens (Aspergillus flavus, A. fumigatus, C. albicans, Campylobacter spp., Salmonella spp., E. coli, S. aureus, and Enterococcus faecalis). All plant extracts showed effective biofilm inhibition of approximately >50% against the tested micro-organisms [160]. Fathi et al. determined the antibiofilm potential of Malva sylvestris methanolic extract against some human pathogens, such as E. coli, S. aureus, P. aeruginosa, E. faecalis, and K. pneumoniae. The highest biofilm inhibition was examined against S. aureus (89.19%), K. pneumoniae (95.46%), and E. faecalis (98.79%) with 40 μg/mL MIC [161].
Priyanto et al. studied the antibiofilm potential of leaf extract of Paederia foetida against E. coli, Mycobacterium smegmatis with 30–50% inhibition, respectively [162]. Panjaitan et al. evaluated the in vitro antibiofilm potential of ethanol extract of Cinnamomum buramanii against periodontal pathogens, such as Aggregatibacter actinomycetemcomitans and Porphyromonas gingivalis. The outcomes revealed that all C. buramanii concentrations showed effective antibiofilm activity against both periodontal pathogens [163].
Plescia et al. determined the antibiofilm potential of Artemisia arborescens plant extracts against S. aureus (ATCC 25923), E. coli (ATCC 25922), P. aeruginosa (ATCC 15442), E. faecalis (29212), and C. albicans (10231). The hot methanol extract showed the highest antibiofilm activity against S. aureus, E. coli, and C. albicans with 58–67% inhibition [164]. Rhimi et al. investigated the in vitro antibiofilm activity of EOs of Cymbopogon spp. (Cymbopogon proximus and Cymbopogon citratus) against Malassezia furfur and Candida spp. The EOs of C. proximus and C. citratus showed significant biofilm inhibition ranging from 27.65 ± 11.7 to 96.39 ± 2.8 against all the tested organisms. Based on the reported results, the EOs of both Cymbopogon spp. can be used for the prevention of Malassezia and Candida infections [165].
Nazzaro et al. determined the antibiofilm activity of EOs of aerial parts and bulbs of two different cultivars of Allium sativum (Bianco del Veneto, Staravec) against nosocomial and food pathogens S. aureus, E. coli, L. monocytogenes, and Acinetobacter baumannii. The EOs from the bulbs and aerial parts of Bianco del Veneto showed significant inhibitory activity against all tested bacteria, more prominently against L. monocytogenes 64.29–60.55%, respectively. The EOs from the aerial parts of Staravec exhibited effective inhibition more effectively against Acinetobacter baumannii (45.61%), while EOs from the bulbs of Staravec showed no inhibition. The outcomes revealed their potential application as potential antibiofilm agents in the food industry and health sector as well [166].
Gamal El-Din et al. investigated the in vitro antibiofilm potential of EOs of three species of Jatropha flowering plant. The EOs were obtained from J. intigrimma, J. gossypiifolia, and J. roseae. The 7.81, 15.63, 31.25, 62.5, 125, 250, 500, and 1000 μg/mL concentrations were used to evaluate the antibiofilm activity. J. intigrimma EO exhibited 100% biofilm inhibitory activity at 31.25 μg/mL. J. roseae EO showed 100% inhibition at 250 μg/mL, while J. gossypiifolia EO revealed less effective biofilm inhibition even at 1000 μg/mL. However, it can be suggested that J. intigrimma and J. roseae EOs can be used as promising antibiofilms and furthering in vivo investigations is also highly recommended [167]. Djebilli et al. determined the composition profile, antioxidant, and antibiofilm efficacy of EOs from Algerian aromatic plants, including Thymus algeriensis, Eucalyptus globulus, and Origanum glandulosum. The EOs from all three plants showed significant antibiofilm activity with low minimum inhibitory concentrations (MICs) ranging between 0.078–1.25 μg/mL against C. albicans, E. coli, E. faecalis, L. monocytogenes, S. aureus, and S. Typhimurium. The order of biofilm inhibition against the tested bacteria was revealed as T. algeriensis > O. glandulosum > E. globulus EOs. The T. algeriensis EO showed the highest inhibition against L. monocytogenes (80.95%), and E. coli (77.83%) at MICs [168]. Some other biofilm inhibiting plant extracts and essential oils are listed in Table 6.
Bee products are effectively being studied for their wide range of antibacterial, antioxidant, antiviral, antifungal, and anticancerous activities. Honey and its bioactive components are well recognized for their potential antibacterial effects against a wide range of bacteria and even against several antibiotic-resistant bacteria [185,186,187].
Bouchelaghem et al. collected propolis from six Hungarian regions and evaluated the in vitro antibiofilm activity by using ethanolic extract (EEP) alone and in combination with vancomycin against MSSA and MRSA. The EEP significantly prevented planktonic growth. The EEP in combination with vancomycin synergistically showed effective inhibition and degradation against biofilm formation and maturation, respectively. The EEP at a concentration of 200 μg/mL against MSSA and MRSA showed 47–87% biofilm degradation, respectively [188]. Alandejani et al. have determined the antibiofilm activity of four different kinds of honey: Manuka honey (from New Zealand), Buckwheat and Canadian clover honey (from Canada), and Sidr honey (from Yemen). All these honey samples have shown considerable bactericidal activity against P. aeruginosa, methicillin-resistant S. aureus (MRSA), and methicillin-sensitive S. aureus (MSSA) biofilms. Manuka and Sidr honey have notably more effective antibiofilm activities against P. aeruginosa, MSSA, and MRSA, ranging from 91%, 63–82%, and 63–73%, respectively [189]. Manuka honey has also shown effective results against some Gram-negative bacteria, including extended-spectrum β-lactamase (ESBL) and carbapenemase-producing K. pneumoniae [190,191,192], ESBL-producing E. coli [191,193], multidrug-resistant (MDR) P. aeruginosa [191], antibiotic-resistant Ureaplasma urealyticum, and Ureaplasma parvum [194].
Fadl et al. reported the antibiofilm potential of bee venom against biofilm-forming MDR bacteria, such as S. aureus, vancomycin-resistant S. aureus (VRSA), P. aeruginosa, Enterobacter cloacae, and S. haemolyticus. Bee venom showed a considerable reduction in biofilm formation, ranging between 63.8–92% [195]. Bouchelaghem et al. determined the anti-biofilm impact of Hungarian propolis. The ethanolic extract of propolis (EEP) was used to evaluate the antibiofilm effect against MSSA and MRSA by applying a crystal violet assay. The EEP alone and in combination with vancomycin were tested against MSSA and MRSA. The EEP significantly inhibited biofilm development and degraded MSSA and MRSA mature biofilms. The EEP, combined with vancomycin, synergistically enhanced the antibiofilm activity against MRSA [188].

8.2. Nanotechnology and Phyto-Nanotechnology

Nanotechnology-based nanomaterials (NMs) are very small in size (˂100 nm) with a large surface area and may provide several biological, chemical, and biomedical applications [196]. NMs can easily enter into the outer membrane (EPS) to release the antimicrobials to targeted sites without damage. Several NM types have been designed and evaluated to inhibit or eradicate microbial biofilms. Nanoparticles are mainly categorized into two categories: organic nanoparticles (NPs) (including polymers, cyclodextrins (CDs), liposomes, dendrimers, and solid lipid NPs) and inorganic NPs (including metal oxides, quantum dots, metallic NPs, and fullerene) [197]. NMs provide potential microbicidal activity alone or in combination with encapsulated drugs [198]. The NMs are reported to provide a promising therapeutic potential for developing significant antibiofilm action [199]. The development of plant-derived nanoparticles (NPs) has emerged as an innovative approach in the era of nanotechnology by synthesizing environmentally friendly substances with little to no toxicity [200]. Several researchers have significantly reported the green synthesis of NPs and their antimicrobial potential against different bacterial biofilms.
Swidan et al. investigated the biofilm inhibition of Ag NPs against enterococcal clinical isolates of the urinary tract (biofilm producing). Three types of Ag NPs were prepared to investigate the antibiofilm activity, including cinnamon Ag NPs synthesized by Cinnamon cassia, ginger Ag NPs synthesized by Zingiber officinale, and the chemically synthesized Ag NPs. The outcomes demonstrated that the chemical and ginger Ag NPs decreased the biofilm formation to 65.32% and 39.14%, and the adhesion to the catheter surface to 69.84% and 42.73%, respectively, and the cinnamon Ag NPs were not as significant. The ginger Ag NPs showed the most effective antibacterial and antiadhesion effects against the enterococcal clinical isolates that can also produce biofilms [201]. Muthulakshmi et al. synthesized the Ag NPs using Terminalia catappa plant leaves and evaluated the in vitro and in vivo antibiofilm potential against the foodborne pathogen L. monocytogenes. The in vitro analysis showed 33–45.5% biofilm inhibition at 50–100 μg/mL, respectively. The in vivo evaluation of Ag NPs using Caenorhabditis Elegans revealed 90% antiadherent activity against L. monocytogenes [202]. Salem et al. synthesized and characterized selenium NPs with orange peel. The biosynthesized Se NPs were used to evaluate the antibiofilm potential against S. aureus, K. pneumonia, and P. aeruginosa. The Se NPs at 0.5 μg/mL concentration showed 95, 88, and 75.5% biofilm inhibition against K. pneumonia, S. aureus, and P. aeruginosa, respectively [203]. Some plant-based NPs, along with their biofilm inhibitory actions, are summarized in Table 7.

8.2.1. Liposomes

Liposomes can significantly enhance the interaction with bacterial membranes and mediate penetration into mature biofilms due to their high biocompatibility. The fusogenic liposome can be significantly employed to encapsulate drugs or antimicrobial agents with optimized release to increase the antibiofilm efficacy (Figure 5). Vancomycin encapsulated with fusogenic liposomes (able to fuse with bacterial membranes) enhanced the bactericidal efficacy against S. aureus biofilms [213]. Meropenem-loaded nanoliposome with a dosage of ≥1.5 µg/mL entirely eradicated the P. aeruginosa biofilm [214]. The drug encapsulated in liposomes showed more effective results than the pristine drug. Liposomes as nanocarriers allow optimized drug release and inhibit biofilm growth.

8.2.2. Solid Lipid NPs

Solid lipid NPs (SLNs) containing surfactants stabilized lipid cores are spherical colloidal NMs (with 10–1000 nm diameters) [215]. Cefuroxime axetil-containing SLNs (CA-SLNs) were designed and analyzed against S. aureus biofilms. CA-SLNs showed a two-fold higher anti-biofilm activity against S. aureus biofilm. Ninety-seven percent of the free CA was released in 2 h, while it takes 12 h to reach 96% release when it is in the encapsulated form [216].

8.2.3. QS Inhibiting NMs

Numerous QS inhibitor (QSI) NMs have been developed to inhibit or eradicate biofilm formation [217]. These QSI-NMs have numerous benefits over conventional QSIs. NMs can penetrate the biofilm matrix due to their smaller size and high solubility, providing optimized and targeted drug release. A QSI entrapped inside solid-lipid NPs with an ultra-small size (<100 nm diameter) exhibits a 7-fold higher anti-biofilm activity against P. aeruginosa than a free QSI [217]. Tellurium (Te) and selenium (Se) NPs have been designed and tested against P. aeruginosa biofilms. Se-NPs and Te-NPs disrupted QS signaling and significantly reduced the bacterial biovolume, resulting in biofilm formation inhibition of 70–80% [218]. However, they were observed to be less effective in removing mature biofilms.

8.2.4. Cyclodextrins

Nguyen et al. investigated the anti-biofilm activity of a ubiquitous flavanone, naringenin (NAR), encapsulated β-CD and chitosan against the biofilm of E. coli [219]. NAR nanoencapsulation showed an effective antibiofilm potential against E. coli. Miconazole encapsulated CDs attached to polypropylene and polyethylene showed 87–96% inhibition against the biofilm formation of C. albicans [220]. CDs coated with drugs covalently bind to the surfaces and effectively inhibit C. albicans biofilm. Thymol and anidulafungin encapsulated cyclodextrins showed significant antibiofilm activity against the biofilms of C. albicans with 75–64% inhibition, respectively [221].

8.2.5. Hydrogels

Many hydrogels have been developed to inhibit or eradicate several microbial biofilms. Some of these, such as metal nanoparticle-based hydrogels [222,223], chitosan-based hydrogels [224,225], bovine serum albumin (BSA) protein-based hydrogels [226], and pentapeptide-based supramolecular hydrogels [227], have exhibited significant biofilm inhibitory and eradication activities against several bacterial biofilms [228]. Some bacteriophages [229,230], drugs [231], and different antimicrobials or active compounds [232] have also been encapsulated into hydrogels, providing efficient anti-biofilm activities with the optimized release of active compounds against different multidrug-resistant bacteria, such as P. aeruginosa, S. aureus, Acinetobacter baumannii, etc. Hydrogels can be used as promising biomaterials to treat and control multidrug-resistant bacterial infections. Some other NMs have also been reported for combating bacterial biofilms (Table 8).

8.3. Microbes and Marine-Derived Anti-Biofilm Compounds

Many micro-organisms produce several types of bioactive molecules with anti-microbial properties to benefit from other micro-organisms. Several investigations have determined the various secondary metabolites with potential anti-biofilm properties extracted from different bacterial and fungal species. Streptomyces species are known as the most promising sources of biofilm-controlling compounds. Methanolic compounds extracted from Streptomyces sp. (strain MUC 125) showed potential anti-biofilm activity against MRSA due to its 2,3-dihydroxybenzoic acid-mediated iron chelating ability [260]. Ethyl-acetate secondary metabolites extracted from Streptomyces sp. (isolated from Iraqi marine sediment) exhibited significant potential for designing and developing new anti-biotics for treating urinary tract infections. The extract showed biofilm inhibition against P. mirabilis (uropathogenic bacteria) even at sub-MIC by interrupting the QS signals [261].
Marine-derived bacteria are well categorized as a prime microbial group found in the marine environment. Peach et al. discovered and characterized the auromomycin chromophore as a potential inhibitor against V. cholera biofilms at 60.1 µM (IC50). Additionally, the inhibitory effect of auromomycin was significantly enhanced by adding some antibiotics (such as tetracycline, ciprofloxacin, and ceftazidime) at their sub-inhibitory concentrations [262]. A red algae-halogenated furanone, Dilsea pulchra, has been reported to have effective anti-biofilm properties [263]. Some synthetic and natural brominated furanones have been reported as significant QS inhibitors against gram-negative and positive bacterial species [264,265,266,267].
Pereira et al. determined the biofilm inhibitory effect of brominated alkylidene lactams (compounds 16, Figure 6) against S. mutans, S. epidermidis, S. aureus, and P. aeruginosa biofilms [268]. The compound-1, γ-hydroxy-γ-lactam, and the compound-2, (E)- γ-alkylidene- γ-lactam were found to be most effective against S. epidermidis with IC50 values of 13.3 and 12.2 µg/mL, respectively. The compound-3 and 4, and (Z)-γ-alkylidene-γ-lactam were observed to be most significant against P. aeruginosa with IC50 values of 0.7 and 0.6 µg/mL, respectively. The compound-5 was most effective against S. aureus with 53.1% biofilm inhibition at 44 µg/mL. The compound-2, 4, and 6 inhibited the biofilm formation of S. mutans [268]. Antibiofilm activity of natural and synthetic cembranoid compounds has been analyzed against P. aeruginosa, V. harveyi, S. aureus, and Chromobacterium violaceum [269].
Actinobacteria are gram-positive and the most versatile bacteria in nature, ranging from aerobic_anaerobic, motile_nonmotile, and sporing_nonsporing bacteria [270]. Actinobacterial spp. produces several secondary metabolites that can act as antibacterial, antiviral, antifungal, and anticancer agents. Some of the recently reported antibiofilm actinobacteria are listed in Table 9. Song et al. revealed that some bacterial strains derived from coral Pocillopora damicornis showed anti-biofilm activity by QS inhibition. Predominantly, H12-Vibrio alginolyticus (a coral symbiotic bacteria) inhibited the biofilm formation of P. aeruginosa PAO1 by interfering with rhl and the las system [271].
Chen et al. isolated three bioactive compounds (benzyl benzoate, 2-methyl-N-(2’-phenylethyl)-butyramide, and 2-methyl-N-(2’-phenylethyl)-butyramide) from a marine bacterium Oceanobacillus sp. (XC22919) and analyzed their antibiofilm activity. All three compounds exhibited significant biofilm inhibitory activity against P. aeruginosa biofilm in a dose-dependent manner by inhibiting QS activity [272].

9. Methodology

Data Collection Criteria

A bibliographic search was conducted to screen relevant scientific publications before August 2022, on bacterial biofilm formation, resistance development mechanisms, and biofilm control strategies. The online search engines of Google scholar, PubMed, ScienceDirect, and Web of Science Core Collection databases were used. The search strategy consisted of separate or simultaneous use, under different combination forms, of several particular keywords including “biofilm formation”, “microbial biofilm composition”, antimicrobial resistance”, “plant-derived anti-biofilm compounds”, “bee products”, “marine derived compounds”, and “phyto-nanotechnology”. Firstly, all titles and abstracts of the search results were individually examined to evaluate whether the articles met inclusion criteria, meaning reporting results in evaluating the antibiofilm activity of plant products, bee products, and nanomaterials against biofilm-forming bacteria with significant outcomes. All the duplicated papers and those published in a language other than English or irrelevant publications that did not contribute to retrieving meaningful results in the goal of assessing natural strategies as potential weapons against bacterial biofilms were excluded. Furthermore, the selected articles were entirely read to obtain significant material based on their experimental outcomes.

10. Conclusions

Over the last three decades, biofilm formation has become a potential threat in the food and health sector. Many biofilm-forming microbial species can develop resistance to harsh environmental conditions. Several antibiotics and different disinfectants are used in hospitals and the food industry. Several biofilm-forming foodborne pathogens have been found to cause outbreaks. Several chronic infections are associated with biofilm-forming microbes. Therefore, several strategies have been designed and analyzed to inhibit microbial growth. However, the emergence of antimicrobial and multidrug resistance forced researchers to study all growth features of microbes and their resistance mechanisms for developing effective combating strategies and significant biofilm-controlling compounds. Many researchers are discovering and developing effective anti-biofilm compounds using natural products. In this review, recent studies were reviewed, focusing on biofilm-controlling compounds, such as natural plants, bee products, nanomaterials, microbes, and marine-derived anti-biofilm compounds to reduce or eradicate microbial biofilms and their associated infections as well. These antibiofilm compounds possess a significant potential to overcome the antimicrobial resistance linked to biofilm formation. On the basis of their potential antibiofilm properties with less to no toxicity and increased bioavailability, they can be suggested to treat biofilm-associated infections.

Author Contributions

Conceptualization and manuscript preparation, writing—review and editing, S.T.A.; literature review and draft preparation, S.R.A.S., D.A.-A., A.A., V.H. and K.Z.; writing—review and editing, S.Z.H., F.R.I., K.M. and D.M.; visualization, M.I., A.M. and C.A.; supervision, writing—review and editing, K.I. and U.A. All authors have read and agreed to the published version of the manuscript.

Funding

This review is supported by the project “Increasing the impact of excellence research on the capacity for innovation and technology transfer within USAMVB Timișoara” code 6PFE, submitted in the competition Program 1—development of the national system of research—development, Subprogram 1.2—institutional performance, institutional development projects—development projects of excellence in RDI.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declared no conflicts of interest.

References

  1. Percival, S.L.; Malic, S.; Cruz, H.; Williams, D.W. Introduction to biofilms. In Biofilms and Veterinary Medicine; Springer: Berlin/Heidelberg, Germany, 2011; pp. 41–68. [Google Scholar]
  2. Asma, S.T.; Imre, K.; Morar, A.; Herman, V.; Acaroz, U.; Mukhtar, H.; Arslan-Acaroz, D.; Shah, S.R.A.; Gerlach, R. An overview of biofilm formation–combating strategies and mechanisms of action of antibiofilm agents. Life 2022, 12, 1110. [Google Scholar] [CrossRef] [PubMed]
  3. Burmølle, M.; Ren, D.; Bjarnsholt, T.; Sørensen, S.J. Interactions in multispecies biofilms: Do they actually matter? Trends Microbiol. 2014, 22, 84–91. [Google Scholar] [CrossRef] [PubMed]
  4. Xue, Z.; Sendamangalam, V.R.; Gruden, C.L.; Seo, Y. Multiple roles of extracellular polymeric substances on resistance of biofilm and detached clusters. ES&T 2012, 46, 13212–13219. [Google Scholar]
  5. Del Pozo, J.L. Biofilm-related disease. Expert Rev. Anti-Infect. 2018, 16, 51–65. [Google Scholar] [CrossRef] [PubMed]
  6. 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] [PubMed]
  7. Davies, J.; Davies, D. Origins and evolution of antibiotic resistance. Microbiol. Mol. Biol. Rev. 2010, 74, 417–433. [Google Scholar] [CrossRef] [Green Version]
  8. Alam, M.M.; Islam, M.; Wahab, A.; Billah, M. Antimicrobial resistance crisis and combating approaches. J. Med. 2019, 20, 38–45. [Google Scholar] [CrossRef]
  9. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Rasool, M.H.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U. Antibiotic resistance: A rundown of a global crisis. Infect. Drug Resist. 2018, 11, 1645. [Google Scholar] [CrossRef] [Green Version]
  10. Sinclair, J.R. Importance of a One Health approach in advancing global health security and the Sustainable Development Goals. Rev. Sci. Tech. Off. Int. Épizoot. 2019, 38, 145–154. [Google Scholar] [CrossRef]
  11. Karygianni, L.; Ren, Z.; Koo, H.; Thurnheer, T. Biofilm matrixome: Extracellular components in structured microbial communities. Trends Microbiol. 2020, 28, 668–681. [Google Scholar] [CrossRef]
  12. Muhammad, M.H.; Idris, A.L.; Fan, X.; Guo, Y.; Yu, Y.; Jin, X.; Qiu, J.; Guan, X.; Huang, T. Beyond risk: Bacterial biofilms and their regulating approaches. Front. Microbiol. 2020, 11, 928. [Google Scholar] [CrossRef] [PubMed]
  13. Kimkes, T.E.; Heinemann, M. How bacteria recognise and respond to surface contact. FEMS Microbiol. Rev. 2020, 44, 106–122. [Google Scholar] [CrossRef]
  14. Aparna, M.S.; Yadav, S. Biofilms: Microbes and disease. Braz. J. Infect. Dis. 2008, 12, 526–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lutz, J.K.; Lee, J. Prevalence and antimicrobial-resistance of Pseudomonas aeruginosa in swimming pools and hot tubs. Int. J. Environ. Res. Public Health 2011, 8, 554–564. [Google Scholar] [CrossRef] [Green Version]
  16. Beloin, C.; Houry, A.; Froment, M.; Ghigo, J.-M.; Henry, N. A short–time scale colloidal system reveals early bacterial adhesion dynamics. PLoS Biol. 2008, 6, e167. [Google Scholar] [CrossRef] [PubMed]
  17. Dunne, W.M., Jr. Bacterial adhesion: Seen any good biofilms lately? Clin. Microbiol. Rev. 2002, 15, 155–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. 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]
  19. Kokare, C.; Chakraborty, S.; Khopade, A.; Mahadik, K.R. Biofilm: Importance and applications. Indian J. Biotechnol. 2009, 8, 159–168. [Google Scholar]
  20. Sauer, K.; Camper, A.K.; Ehrlich, G.D.; Costerton, J.W.; Davies, D.G. Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J. Bacteriol. 2002, 184, 1140–1154. [Google Scholar] [CrossRef] [Green Version]
  21. Lu, T.K.; Collins, J.J. Dispersing biofilms with engineered enzymatic bacteriophage. Proc. Natl. Acad. Sci. USA 2007, 104, 11197–11202. [Google Scholar] [CrossRef] [Green Version]
  22. Evans, L.V. Biofilms: Recent Advances in Their Study and Control, 1st ed.; CRC Press: London, UK, 2000. [Google Scholar]
  23. Flemming, H.-C.; Wingender, J.; Griegbe, T.; Mayer, C. Physico-chemical properties of biofilms. In Biofilms: Recent Advances in Their Study and Control; Harwood Academic Publishers: Amsterdam, The Netherlands, 2000; pp. 19–34. [Google Scholar]
  24. Flemming, H.; Wingender, J. The biofilm matrix. Nat. Publ. Gr. 2010, 8, 623–633. [Google Scholar] [CrossRef] [PubMed]
  25. Lembre, P.; Lorentz, C.; Di Martino, P. Exopolysaccharides of the biofilm matrix: A complex biophysical world. In The Complex World of Polysaccharides; Karunaratne, D., Ed.; IntechOpen: London, UK, 2012. [Google Scholar]
  26. Limoli, D.H.; Jones, C.J.; Wozniak, D.J. Bacterial extracellular polysaccharides in biofilm formation and function. Microbiol Spectr. 2015, 3, 10.1128. [Google Scholar] [CrossRef] [PubMed]
  27. Rehm, B.H. Bacterial polymers: Biosynthesis, modifications and applications. Nat. Rev. Microbiol. 2010, 8, 578–592. [Google Scholar] [CrossRef] [PubMed]
  28. Kaplan, J.B.; Velliyagounder, K.; Ragunath, C.; Rohde, H.; Mack, D.; Knobloch, J.K.-M.; Ramasubbu, N. Genes involved in the synthesis and degradation of matrix polysaccharide in Actinobacillus actinomycetemcomitans and Actinobacillus pleuropneumoniae biofilms. J. Bacteriol. 2004, 186, 8213–8220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Stevenson, G.; Andrianopoulos, K.; Hobbs, M.; Reeves, P.R. Organization of the Escherichia coli K-12 gene cluster responsible for production of the extracellular polysaccharide colanic acid. J. Bacteriol. 1996, 178, 4885–4893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Sayem, S.; Manzo, E.; Ciavatta, L.; Tramice, A.; Cordone, A.; Zanfardino, A.; De Felice, M.; Varcamonti, M. Anti-biofilm activity of an exopolysaccharide from a sponge-associated strain of Bacillus licheniformis. Microb. Cell Factories 2011, 10, 74. [Google Scholar] [CrossRef] [Green Version]
  31. Wingender, J.; Strathmann, M.; Rode, A.; Leis, A.; Flemming, H.-C. [25] Isolation and biochemical characterization of extracellular polymeric substances from Pseudomonas aeruginosa. Meth Enzymol. 2001, 336, 302–314. [Google Scholar]
  32. Ma, L.; Jackson, K.D.; Landry, R.M.; Parsek, M.R.; Wozniak, D.J. Analysis of Pseudomonas aeruginosa conditional psl variants reveals roles for the psl polysaccharide in adhesion and maintaining biofilm structure postattachment. J. Bacteriol. 2006, 188, 8213–8221. [Google Scholar] [CrossRef] [Green Version]
  33. Zogaj, X.; Nimtz, M.; Rohde, M.; Bokranz, W.; Römling, U. The multicellular morphotypes of Salmonella Typhimurium and Escherichia coli produce cellulose as the second component of the extracellular matrix. Mol. Microbiol. 2001, 39, 1452–1463. [Google Scholar] [CrossRef]
  34. Conrad, A.; Kontro, M.; Keinänen, M.M.; Cadoret, A.; Faure, P.; Mansuy-Huault, L.; Block, J.-C. Fatty acids of lipid fractions in extracellular polymeric substances of activated sludge flocs. Lipids 2003, 38, 1093–1105. [Google Scholar] [CrossRef]
  35. Fong, J.; Yildiz, F. Biofilm matrix proteins. Microbiol. Spectr. 2015, 3, 10.1128. [Google Scholar] [CrossRef] [Green Version]
  36. Kaplan, J.B.; Ragunath, C.; Ramasubbu, N.; Fine, D.H. Detachment of Actinobacillus actinomycetemcomitans biofilm cells by an endogenous β-hexosaminidase activity. J. Bacteriol. 2003, 185, 4693–4698. [Google Scholar] [CrossRef] [Green Version]
  37. Martí, M.; Trotonda, M.P.; Tormo-Más, M.Á.; Vergara-Irigaray, M.; Cheung, A.L.; Lasa, I.; Penadés, J.R. Extracellular proteases inhibit protein-dependent biofilm formation in Staphylococcus aureus. Microbes Infect. 2010, 12, 55–64. [Google Scholar] [CrossRef]
  38. Nijland, R.; Hall, M.J.; Burgess, J.G. Dispersal of biofilms by secreted, matrix degrading, bacterial DNase. PLoS ONE 2010, 5, e15668. [Google Scholar] [CrossRef] [Green Version]
  39. Toyofuku, M.; Roschitzki, B.; Riedel, K.; Eberl, L. Identification of proteins associated with the Pseudomonas aeruginosa biofilm extracellular matrix. J. Proteome Res. 2012, 11, 4906–4915. [Google Scholar] [CrossRef]
  40. Wingender, J.; Jaeger, K.-E.; Flemming, H.-C. Interaction between extracellular polysaccharides and enzymes. In Microbial Extracellular Polymeric Substances; Springer: Berlin/Heidelberg, Germany, 1999; pp. 231–251. [Google Scholar]
  41. Izano, E.A.; Amarante, M.A.; Kher, W.B.; Kaplan, J.B. Differential roles of poly-N-acetylglucosamine surface polysaccharide and extracellular DNA in Staphylococcus aureus and Staphylococcus epidermidis biofilms. Appl. Environ. Microbiol. 2008, 74, 470–476. [Google Scholar] [CrossRef] [Green Version]
  42. Wingender, J.; Neu, T.R.; Flemming, H.C. What are bacterial extracellular polymeric substances? In Microbial Extracellular Polymeric Substances; Springer: Berlin/Heidelberg, Germany, 1999; pp. 1–19. [Google Scholar]
  43. Molin, S.; Tolker-Nielsen, T. Gene transfer occurs with enhanced efficiency in biofilms and induces enhanced stabilisation of the biofilm structure. Curr. Opin. Biotechnol. 2003, 14, 255–261. [Google Scholar] [CrossRef]
  44. Yang, L.; Barken, K.B.; Skindersoe, M.E.; Christensen, A.B.; Givskov, M.; Tolker-Nielsen, T. Effects of iron on DNA release and biofilm development by Pseudomonas aeruginosa. Microbiology 2007, 153, 1318–1328. [Google Scholar] [CrossRef] [Green Version]
  45. Whitchurch, C.B.; Tolker-Nielsen, T.; Ragas, P.C.; Mattick, J.S. Extracellular DNA required for bacterial biofilm formation. Science 2002, 295, 1487. [Google Scholar] [CrossRef]
  46. Vilain, S.; Pretorius, J.M.; Theron, J.; Brözel, V.S. DNA as an adhesin: Bacillus cereus requires extracellular DNA to form biofilms. Appl. Environ. Microbiol. 2009, 75, 2861–2868. [Google Scholar] [CrossRef] [Green Version]
  47. Okshevsky, M.; Meyer, R.L. The role of extracellular DNA in the establishment, maintenance and perpetuation of bacterial biofilms. Crit. Rev. Microbiol. 2015, 41, 341–352. [Google Scholar] [CrossRef] [PubMed]
  48. Harmsen, M.; Lappann, M.; Knøchel, S.; Molin, S. Role of extracellular DNA during biofilm formation by Listeria monocytogenes. Appl. Environ. Microbiol. 2010, 76, 2271–2279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wilton, M.; Charron-Mazenod, L.; Moore, R.; Lewenza, S. Extracellular DNA acidifies biofilms and induces aminoglycoside resistance in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2016, 60, 544–553. [Google Scholar] [CrossRef]
  50. Neu, T.R.; Poralla, K. An amphiphilic polysaccharide from an adhesive Rhodococcus strain. FEMS Microbiol. Lett. 1988, 49, 389–392. [Google Scholar] [CrossRef]
  51. Kjelleberg, S.; Givskov, M. Biofilm Mode of Life; Horizon Bioscience; The University of New South Wales: Sydney, Australia; Technical University of Denmark: Lyngby, Denmark, 2007. [Google Scholar]
  52. Sutherland, I.W. The biofilm matrix–an immobilized but dynamic microbial environment. Trends Microbiol. 2001, 9, 222–227. [Google Scholar] [CrossRef]
  53. Schmitt, J.; Flemming, H.-C. Water binding in biofilms. Water Sci. Technol. 1999, 39, 77–82. [Google Scholar] [CrossRef]
  54. Lewis, K. Riddle of biofilm resistance. Antimicrob. Agents Chemother. 2001, 45, 999–1007. [Google Scholar] [CrossRef] [Green Version]
  55. Rani, S.A.; Pitts, B.; Beyenal, H.; Veluchamy, R.A.; Lewandowski, Z.; Davison, W.M.; Buckingham-Meyer, K.; Stewart, P.S. Spatial patterns of DNA replication, protein synthesis, and oxygen concentration within bacterial biofilms reveal diverse physiological states. J. Bacteriol. 2007, 189, 4223–4233. [Google Scholar] [CrossRef] [Green Version]
  56. Bridier, A.; Dubois-Brissonnet, F.; Boubetra, A.; Thomas, V.; Briandet, R. The biofilm architecture of sixty opportunistic pathogens deciphered using a high throughput CLSM method. J. Microbiol. Methods 2010, 82, 64–70. [Google Scholar] [CrossRef]
  57. Lear, G.; Lewis, G.D. Microbial Biofilms: Current Research and Applications; Lincoln University: Christchurch, New Zealand; University of Auckland: Auckland, New Zealand, 2012. [Google Scholar]
  58. Toyofuku, M.; Inaba, T.; Kiyokawa, T.; Obana, N.; Yawata, Y.; Nomura, N. Environmental factors that shape biofilm formation. Biosci. Biotechnol. Biochem. 2016, 80, 7–12. [Google Scholar] [CrossRef]
  59. Tielen, P.; Strathmann, M.; Jaeger, K.-E.; Flemming, H.-C.; Wingender, J. Alginate acetylation influences initial surface colonization by mucoid Pseudomonas aeruginosa. Microbiol. Res. 2005, 160, 165–176. [Google Scholar] [CrossRef] [PubMed]
  60. Watnick, P.I.; Kolter, R. Steps in the development of a Vibrio cholerae El Tor biofilm. Mol. Microbiol. 1999, 34, 586–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Danese, P.N.; Pratt, L.A.; Kolter, R. Exopolysaccharide production is required for development of Escherichia coli K-12 biofilm architecture. J. Bacteriol. 2000, 182, 3593–3596. [Google Scholar] [CrossRef] [PubMed]
  62. Branda, S.S.; Chu, F.; Kearns, D.B.; Losick, R.; Kolter, R. A major protein component of the Bacillus subtilis biofilm matrix. Mol. Microbiol. 2006, 59, 1229–1238. [Google Scholar] [CrossRef]
  63. Manfiolli, A.O.; Dos Reis, T.F.; de Assis, L.J.; de Castro, P.A.; Silva, L.P.; Hori, J.I.; Walker, L.A.; Munro, C.A.; Rajendran, R.; Ramage, G. Mitogen activated protein kinases (MAPK) and protein phosphatases are involved in Aspergillus fumigatus adhesion and biofilm formation. Cell Surf. 2018, 1, 43–56. [Google Scholar] [CrossRef]
  64. McCourt, J.; O′Halloran, D.P.; McCarthy, H.; O′Gara, J.P.; Geoghegan, J.A. Fibronectin-binding proteins are required for biofilm formation by community-associated methicillin-resistant Staphylococcus aureus strain LAC. FEMS Microbiol. Lett. 2014, 353, 157–164. [Google Scholar] [CrossRef] [Green Version]
  65. Ye, Y.; Ling, N.; Gao, J.; Zhang, X.; Zhang, M.; Tong, L.; Zeng, H.; Zhang, J.; Wu, Q. Roles of outer membrane protein W (OmpW) on survival, morphology, and biofilm formation under NaCl stresses in Cronobacter sakazakii. Int. J. Dairy Sci. 2018, 101, 3844–3850. [Google Scholar] [CrossRef]
  66. Liang, X.; Chen, Y.-Y.M.; Ruiz, T.; Wu, H. New cell surface protein involved in biofilm formation by Streptococcus parasanguinis. Infect. Immun. 2011, 79, 3239–3248. [Google Scholar] [CrossRef] [Green Version]
  67. Camilli, A.; Bassler, B.L. Bacterial small-molecule signaling pathways. Science 2006, 311, 1113–1116. [Google Scholar] [CrossRef] [Green Version]
  68. Kalia, V.C. Quorum sensing inhibitors: An overview. Biotechnol. Adv. 2013, 31, 224–245. [Google Scholar] [CrossRef]
  69. Xue, T.; Ni, J.; Shang, F.; Chen, X.; Zhang, M. Autoinducer-2 increases biofilm formation via an ica-and bhp-dependent manner in Staphylococcus epidermidis RP62A. Microbes Infect. 2015, 17, 345–352. [Google Scholar] [CrossRef] [PubMed]
  70. Lee, K.; Yoon, S.S. Pseudomonas aeruginosa biofilm, a programmed bacterial life for fitness. J. Microbiol. Biotechnol. 2017, 27, 1053–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Mikkelsen, H.; Sivaneson, M.; Filloux, A. Key two-component regulatory systems that control biofilm formation in Pseudomonas aeruginosa. Environ. Microbiol. 2011, 13, 1666–1681. [Google Scholar] [CrossRef] [PubMed]
  72. Shao, Y.; Bassler, B.L. Quorum-sensing non-coding small RNAs use unique pairing regions to differentially control mRNA targets. Mol. Microbiol. 2012, 83, 599–611. [Google Scholar] [CrossRef] [PubMed]
  73. Moreno, R.; Fonseca, P.; Rojo, F. Two small RNAs, CrcY and CrcZ, act in concert to sequester the Crc global regulator in Pseudomonas putida, modulating catabolite repression. Mol. Microbiol. 2012, 83, 24–40. [Google Scholar] [CrossRef]
  74. Jørgensen, M.G.; Nielsen, J.S.; Boysen, A.; Franch, T.; Møller-Jensen, J.; Valentin-Hansen, P. Small regulatory RNAs control the multi-cellular adhesive lifestyle of Escherichia coli. Mol. Microbiol. 2012, 84, 36–50. [Google Scholar] [CrossRef] [Green Version]
  75. Thomason, M.K.; Fontaine, F.; De Lay, N.; Storz, G. A small RNA that regulates motility and biofilm formation in response to changes in nutrient availability in Escherichia coli. Mol. Microbiol. 2012, 84, 17–35. [Google Scholar] [CrossRef]
  76. Holmqvist, E.; Reimegård, J.; Sterk, M.; Grantcharova, N.; Römling, U.; Wagner, E.G.H. Two antisense RNAs target the transcriptional regulator CsgD to inhibit curli synthesis. EMBO J. 2010, 29, 1840–1850. [Google Scholar] [CrossRef] [Green Version]
  77. Monteiro, C.; Papenfort, K.; Hentrich, K.; Ahmad, I.; Le Guyon, S.; Reimann, R.; Grantcharova, N.; Römling, U. Hfq and Hfq-dependent small RNAs are major contributors to multicellular development in Salmonella enterica serovar Typhimurium. RNA Biol. 2012, 9, 489–502. [Google Scholar] [CrossRef] [Green Version]
  78. Mika, F.; Busse, S.; Possling, A.; Berkholz, J.; Tschowri, N.; Sommerfeldt, N.; Pruteanu, M.; Hengge, R. Targeting of csgD by the small regulatory RNA RprA links stationary phase, biofilm formation and cell envelope stress in Escherichia coli. Mol. Microbiol. 2012, 84, 51–65. [Google Scholar] [CrossRef] [Green Version]
  79. Papenfort, K.; Said, N.; Welsink, T.; Lucchini, S.; Hinton, J.C.; Vogel, J. Specific and pleiotropic patterns of mRNA regulation by ArcZ, a conserved, Hfq-dependent small RNA. Mol. Microbiol. 2009, 74, 139–158. [Google Scholar] [CrossRef] [PubMed]
  80. Mandin, P.; Gottesman, S. Integrating anaerobic/aerobic sensing and the general stress response through the ArcZ small RNA. EMBO J. 2010, 29, 3094–3107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Jackson, D.W.; Suzuki, K.; Oakford, L.; Simecka, J.W.; Hart, M.E.; Romeo, T. Biofilm formation and dispersal under the influence of the global regulator CsrA of Escherichia coli. J. Bacteriol. 2002, 184, 290–301. [Google Scholar] [CrossRef] [PubMed]
  82. Heroven, A.K.; Böhme, K.; Rohde, M.; Dersch, P. A Csr-type regulatory system, including small non-coding RNAs, regulates the global virulence regulator RovA of Yersinia pseudotuberculosis through RovM. Mol. Microbiol. 2008, 68, 1179–1195. [Google Scholar] [CrossRef] [PubMed]
  83. Sonnleitner, E.; Gonzalez, N.; Sorger-Domenigg, T.; Heeb, S.; Richter, A.S.; Backofen, R.; Williams, P.; Hüttenhofer, A.; Haas, D.; Bläsi, U. The small RNA PhrS stimulates synthesis of the Pseudomonas aeruginosa quinolone signal. Mol. Microbiol. 2011, 80, 868–885. [Google Scholar] [CrossRef] [PubMed]
  84. Majdalani, N.; Chen, S.; Murrow, J.; St John, K.; Gottesman, S. Regulation of RpoS by a novel small RNA: The characterization of RprA. Mol. Microbiol. 2001, 39, 1382–1394. [Google Scholar] [CrossRef]
  85. Yin, W.; Wang, Y.; Liu, L.; He, J. Biofilms: The microbial “protective clothing” in extreme environments. Int. J. Mol. Sci. 2019, 20, 3423. [Google Scholar] [CrossRef] [Green Version]
  86. Hathroubi, S.; Mekni, M.A.; Domenico, P.; Nguyen, D.; Jacques, M. Biofilms: Microbial shelters against antibiotics. Microb. Drug Resist. 2017, 23, 147–156. [Google Scholar] [CrossRef]
  87. Tseng, B.S.; Zhang, W.; Harrison, J.J.; Quach, T.P.; Song, J.L.; Penterman, J.; Singh, P.K.; Chopp, D.L.; Packman, A.I.; Parsek, M.R. The extracellular matrix protects Pseudomonas aeruginosa biofilms by limiting the penetration of tobramycin. Environ. Microbiol. 2013, 15, 2865–2878. [Google Scholar]
  88. Zhou, G.; Shi, Q.-S.; Huang, X.-M.; Xie, X.-B. The three bacterial lines of defense against antimicrobial agents. Int. J. Mol. Sci. 2015, 16, 21711–21733. [Google Scholar] [CrossRef] [Green Version]
  89. Stewart, P.S.; Zhang, T.; Xu, R.; Pitts, B.; Walters, M.C.; Roe, F.; Kikhney, J.; Moter, A. Reaction–diffusion theory explains hypoxia and heterogeneous growth within microbial biofilms associated with chronic infections. NPJ Biofilms Microbiomes 2016, 2, 16012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Miao, Y.; Zhou, J.; Chen, C.; Shen, D.; Song, W.; Feng, Y. In vitro adsorption revealing an apparent strong interaction between endophyte Pantoea agglomerans YS19 and host rice. Curr. Microbiol. 2008, 57, 547–551. [Google Scholar] [CrossRef] [PubMed]
  91. Li, Q.; Miao, Y.; Yi, T.; Zhou, J.; Lu, Z.; Feng, Y. SPM43. 1 contributes to acid-resistance of non-symplasmata-forming cells in Pantoea agglomerans YS19. Curr. Microbiol. 2012, 64, 214–221. [Google Scholar] [CrossRef] [PubMed]
  92. Zhao, X.; Yu, Z.; Ding, T. Quorum-sensing regulation of antimicrobial resistance in bacteria. Microorganisms 2020, 8, 425. [Google Scholar] [CrossRef] [PubMed]
  93. Passos da Silva, D.; Schofield, M.C.; Parsek, M.R.; Tseng, B.S. An update on the sociomicrobiology of quorum sensing in gram-negative biofilm development. Pathogens 2017, 6, 51. [Google Scholar] [CrossRef]
  94. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef]
  95. Whiteley, M.; Diggle, S.P.; Greenberg, E.P. Progress in and promise of bacterial quorum sensing research. Nature 2017, 551, 313–320. [Google Scholar] [CrossRef] [Green Version]
  96. Wolska, K.I.; Grudniak, A.M.; Rudnicka, Z.; Markowska, K. Genetic control of bacterial biofilms. J. Appl. Genet. 2016, 57, 225–238. [Google Scholar] [CrossRef] [Green Version]
  97. Pena, R.T.; Blasco, L.; Ambroa, A.; González-Pedrajo, B.; Fernández-García, L.; López, M.; Bleriot, I.; Bou, G.; García-Contreras, R.; Wood, T.K. Relationship between quorum sensing and secretion systems. Front. Microbiol. 2019, 10, 1100. [Google Scholar] [CrossRef] [Green Version]
  98. Blanco, P.; Hernando-Amado, S.; Reales-Calderon, J.A.; Corona, F.; Lira, F.; Alcalde-Rico, M.; Bernardini, A.; Sanchez, M.B.; Martinez, J.L. Bacterial multidrug efflux pumps: Much more than antibiotic resistance determinants. Microorganisms 2016, 4, 14. [Google Scholar] [CrossRef] [Green Version]
  99. Blöcher, R.; Rodarte Ramírez, A.; Castro-Escarpulli, G.; Curiel-Quesada, E.; Reyes-Arellano, A. Design, synthesis, and evaluation of alkyl-quinoxalin-2 (1 1H)-one derivatives as anti-quorum sensing molecules, inhibiting biofilm formation in Aeromonas caviae Sch3. Molecules 2018, 23, 3075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Kumar, A.; Schweizer, H.P. Bacterial resistance to antibiotics: Active efflux and reduced uptake. Adv. Drug Deliv. Rev. 2005, 57, 1486–1513. [Google Scholar] [CrossRef] [PubMed]
  101. Singh, S.; Singh, S.K.; Chowdhury, I.; Singh, R. Understanding the mechanism of bacterial biofilms resistance to antimicrobial agents. Open J. Med. Microbiol. 2017, 11, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Van Acker, H.; Coenye, T. The role of efflux and physiological adaptation in biofilm tolerance and resistance. J. Biol. Chem. 2016, 291, 12565–12572. [Google Scholar] [CrossRef]
  103. Kunz Coyne, A.J.; El Ghali, A.; Holger, D.; Rebold, N.; Rybak, M.J. Therapeutic strategies for emerging multidrug-resistant Pseudomonas aeruginosa. Infect Dis Ther. 2022, 11, 661–682. [Google Scholar] [CrossRef]
  104. Zhang, L.; Mah, T.-F. Involvement of a novel efflux system in biofilm-specific resistance to antibiotics. J. Bacteriol. 2008, 190, 4447–4452. [Google Scholar] [CrossRef] [Green Version]
  105. Buroni, S.; Matthijs, N.; Spadaro, F.; Van Acker, H.; Scoffone, V.C.; Pasca, M.R.; Riccardi, G.; Coenye, T. Differential roles of RND efflux pumps in antimicrobial drug resistance of sessile and planktonic Burkholderia cenocepacia cells. Antimicrob. Agents Chemother. 2014, 58, 7424–7429. [Google Scholar] [CrossRef] [Green Version]
  106. Shoji, M.M.; Chen, A.F. Biofilms in periprosthetic joint infections: A review of diagnostic modalities, current treatments, and future directions. J. Knee Surg. 2020, 33, 119–131. [Google Scholar] [CrossRef]
  107. Furner-Pardoe, J.; Anonye, B.O.; Cain, R.; Moat, J.; Ortori, C.A.; Lee, C.; Barrett, D.A.; Corre, C.; Harrison, F. Anti-biofilm efficacy of a medieval treatment for bacterial infection requires the combination of multiple ingredients. Sci. Rep. 2020, 10, 12687. [Google Scholar] [CrossRef]
  108. Vajjala, A.; Biswas, D.; Tay, W.H.; Hanski, E.; Kline, K.A. Streptolysin-induced endoplasmic reticulum stress promotes group A streptococcal host-associated biofilm formation and necrotising fasciitis. Cell. Microbiol. 2019, 21, e12956. [Google Scholar] [CrossRef] [Green Version]
  109. Colombo, A.P.V.; Magalhães, C.B.; Hartenbach, F.A.R.R.; do Souto, R.M.; da Silva-Boghossian, C.M. Periodontal-disease-associated biofilm: A reservoir for pathogens of medical importance. Microb. Pathog. 2016, 94, 27–34. [Google Scholar] [CrossRef] [PubMed]
  110. Roldán, S.; Herrera, D.; Sanz, M. Biofilms and the tongue: Therapeutical approaches for the control of halitosis. Clin. Oral Investig. 2003, 7, 189–197. [Google Scholar] [CrossRef] [PubMed]
  111. Zimmerli, W.; Sendi, P. Orthopaedic biofilm infections. APMIS 2017, 125, 353–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Bakar, M.B.A.; McKimm, J.; Haque, M. Otitis media and biofilm: An overview. Int. J. Nutr. Pharmacol. Neurol. 2018, 8, 70–78. [Google Scholar]
  113. Høiby, N.; Ciofu, O.; Bjarnsholt, T. Pseudomonas aeruginosa biofilms in cystic fibrosis. Future Microbiol. 2010, 5, 1663–1674. [Google Scholar] [CrossRef]
  114. Galie, S.; García-Gutiérrez, C.; Miguélez, E.M.; Villar, C.J.; Lombó, F. Biofilms in the food industry: Health aspects and control methods. Front. Microbiol. 2018, 9, 898. [Google Scholar] [CrossRef]
  115. Lindsay, D.; von Holy, A. What food safety professionals should know about bacterial biofilms. Brit. Food J. 2006, 108, 27–37. [Google Scholar] [CrossRef]
  116. Srey, S.; Jahid, I.K.; Ha, S.-D. Biofilm formation in food industries: A food safety concern. Food Control 2013, 31, 572–585. [Google Scholar] [CrossRef]
  117. Govaert, M.; Smet, C.; Baka, M.; Janssens, T.; Impe, J.V. Influence of incubation conditions on the formation of model biofilms by Listeria monocytogenes and Salmonella Typhimurium on abiotic surfaces. J. Appl. Microbiol. 2018, 125, 1890–1900. [Google Scholar] [CrossRef]
  118. Tang, L.; Pillai, S.; Revsbech, N.P.; Schramm, A.; Bischoff, C.; Meyer, R.L. Biofilm retention on surfaces with variable roughness and hydrophobicity. Biofouling 2011, 27, 111–121. [Google Scholar] [CrossRef]
  119. Makovcova, J.; Babak, V.; Kulich, P.; Masek, J.; Slany, M.; Cincarova, L. Dynamics of mono-and dual-species biofilm formation and interactions between Staphylococcus aureus and Gram-negative bacteria. Microb. Biotechnol. 2017, 10, 819–832. [Google Scholar] [CrossRef] [PubMed]
  120. Van Houdt, R.; Michiels, C. Biofilm formation and the food industry, a focus on the bacterial outer surface. J. Appl. Microbiol. 2010, 109, 1117–1131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Araújo, E.A.; de Andrade, N.J.; da Silva, L.H.M.; de Carvalho, A.F.; de Sá Silva, C.A.; Ramos, A.M. Control of microbial adhesion as a strategy for food and bioprocess technology. Food Bioproc. Tech. 2010, 3, 321–332. [Google Scholar] [CrossRef]
  122. Dhowlaghar, N.; Bansal, M.; Schilling, M.W.; Nannapaneni, R. Scanning electron microscopy of Salmonella biofilms on various food-contact surfaces in catfish mucus. Food Microbiol. 2018, 74, 143–150. [Google Scholar] [CrossRef] [PubMed]
  123. Jindal, S.; Anand, S.; Metzger, L.; Amamcharla, J. A comparison of biofilm development on stainless steel and modified-surface plate heat exchangers during a 17-h milk pasteurization run. Int. J. Dairy Sci. 2018, 101, 2921–2926. [Google Scholar] [CrossRef]
  124. Dutra, T.V.; Fernandes, M.d.S.; Perdoncini, M.R.F.G.; Anjos, M.M.d.; Abreu Filho, B.A.d. Capacity of Escherichia coli and Staphylococcus aureus to produce biofilm on stainless steel surfaces in the presence of food residues. J. Food Process. Preserv. 2018, 42, e13574. [Google Scholar] [CrossRef]
  125. Iñiguez-Moreno, M.; Gutiérrez-Lomelí, M.; Avila-Novoa, M.G. Kinetics of biofilm formation by pathogenic and spoilage microorganisms under conditions that mimic the poultry, meat, and egg processing industries. Int. J. Food Microbiol. 2019, 303, 32–41. [Google Scholar] [CrossRef]
  126. Karaca, B.; Buzrul, S.; Coleri Cihan, A. Anoxybacillus and Geobacillus biofilms in the dairy industry: Effects of surface material, incubation temperature and milk type. Biofouling 2019, 35, 551–560. [Google Scholar] [CrossRef]
  127. Grigore-Gurgu, L.; Bucur, F.I.; Borda, D.; Alexa, E.-A.; Neagu, C.; Nicolau, A.I. Biofilms formed by pathogens in food and food processing environments. Bacterial Biofilms; Dincer, S., Özdenefe, M., Arkut, A., Eds.; Intech Open: London, UK, 2019; pp. 1–32. [Google Scholar]
  128. Kwon, M.; Hussain, M.S.; Oh, D.H. Biofilm formation of Bacillus cereus under food-processing-related conditions. Food Sci. Biotechnol. 2017, 26, 1103–1111. [Google Scholar] [CrossRef]
  129. Klančnik, A.; Šimunović, K.; Sterniša, M.; Ramić, D.; Smole Možina, S.; Bucar, F. Anti-adhesion activity of phytochemicals to prevent Campylobacter jejuni biofilm formation on abiotic surfaces. Phytochem. Rev. 2021, 20, 55–84. [Google Scholar] [CrossRef] [Green Version]
  130. Carrascosa, C.; Raheem, D.; Ramos, F.; Saraiva, A.; Raposo, A. Microbial biofilms in the food industry—A comprehensive review. Int. J. Environ. Res. Public Health. 2021, 18, 2014. [Google Scholar] [CrossRef]
  131. Ricci, E.; Schwinghamer, T.; Fan, D.; Smith, D.L.; Gravel, V. Growth promotion of greenhouse tomatoes with Pseudomonas sp. and Bacillus sp. biofilms and planktonic cells. Appl. Soil Ecol. 2019, 138, 61–68. [Google Scholar] [CrossRef]
  132. Kadariya, J.; Smith, T.C.; Thapaliya, D. Staphylococcus aureus and staphylococcal food-borne disease: An ongoing challenge in public health. Biomed Res. Int. 2014, 2014, 827965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Giaouris, E.; Heir, E.; Desvaux, M.; Hébraud, M.; Møretrø, T.; Langsrud, S.; Doulgeraki, A.; Nychas, G.-J.; Kačániová, M.; Czaczyk, K. Intra-and inter-species interactions within biofilms of important foodborne bacterial pathogens. Front. Microbiol. 2015, 6, 841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hossain, M.I.; Mizan, M.F.R.; Ashrafudoulla, M.; Nahar, S.; Joo, H.-J.; Jahid, I.K.; Park, S.H.; Kim, K.-S.; Ha, S.-D. Inhibitory effects of probiotic potential lactic acid bacteria isolated from kimchi against Listeria monocytogenes biofilm on lettuce, stainless-steel surfaces, and MBEC™ biofilm device. LWT 2020, 118, 108864. [Google Scholar] [CrossRef]
  135. Rodríguez-Saavedra, M.; de Llano, D.G.; Beltran, G.; Torija, M.-J.; Moreno-Arribas, M.V. Pectinatus spp.- unpleasant and recurrent brewing spoilage bacteria. Int. J. Food Microbiol. 2021, 336, 108900. [Google Scholar] [CrossRef] [PubMed]
  136. Smith, A.M.; Tau, N.P.; Smouse, S.L.; Allam, M.; Ismail, A.; Ramalwa, N.R.; Disenyeng, B.; Ngomane, M.; Thomas, J. Outbreak of Listeria monocytogenes in South Africa, 2017–2018: Laboratory activities and experiences associated with whole-genome sequencing analysis of isolates. Foodborne Pathog. Dis. 2019, 16, 524–530. [Google Scholar] [CrossRef] [Green Version]
  137. Bhattacharya, A.; Shantikumar, S.; Beaufoy, D.; Allman, A.; Fenelon, D.; Reynolds, K.; Normington, A.; Afza, M.; Todkill, D. Outbreak of Clostridium perfringens food poisoning linked to leeks in cheese sauce: An unusual source. Epidemiol. Infect. 2020, 148, e43. [Google Scholar] [CrossRef] [Green Version]
  138. Mikhail, A.; Jenkins, C.; Dallman, T.; Inns, T.; Douglas, A.; Martín, A.; Fox, A.; Cleary, P.; Elson, R.; Hawker, J. An outbreak of Shiga toxin-producing Escherichia coli O157: H7 associated with contaminated salad leaves: Epidemiological, genomic and food trace back investigations. Epidemiol. Infect. 2018, 146, 187–196. [Google Scholar] [CrossRef] [Green Version]
  139. Vaughn, E.L.; Vo, Q.T.; Vostok, J.; Stiles, T.; Lang, A.; Brown, C.M.; Klevens, R.M.; Madoff, L. Linking epidemiology and whole-genome sequencing to investigate Salmonella outbreak, Massachusetts, USA, 2018. Emerg. Infect. Dis. 2020, 26, 1538. [Google Scholar] [CrossRef]
  140. Kenyon, J.; Inns, T.; Aird, H.; Swift, C.; Astbury, J.; Forester, E.; Decraene, V. Campylobacter outbreak associated with raw drinking milk, North West England, 2016. Epidemiol. Infect. 2020, 148, e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Denayer, S.; Delbrassinne, L.; Nia, Y.; Botteldoorn, N. Food-borne outbreak investigation and molecular typing: High diversity of Staphylococcus aureus strains and importance of toxin detection. Toxins 2017, 9, 407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Chen, J.; Zhang, R.; Qi, X.; Zhou, B.; Wang, J.; Chen, Y.; Zhang, H. Epidemiology of foodborne disease outbreaks caused by Vibrio parahaemolyticus during 2010–2014 in Zhejiang Province, China. Food Control 2017, 77, 110–115. [Google Scholar] [CrossRef]
  143. Griffiths, M.; Schraft, H. Bacillus cereus food poisoning. In Foodborne diseases; Elsevier: Amsterdam, The Netherlands, 2017; pp. 395–405. [Google Scholar]
  144. Wu, Y.; Wen, J.; Ma, Y.; Ma, X.; Chen, Y. Epidemiology of foodborne disease outbreaks caused by Vibrio parahaemolyticus, China, 2003–2008. Food Control 2014, 46, 197–202. [Google Scholar] [CrossRef]
  145. Popovic, I.; Heron, B.; Covacin, C. Listeria: An Australian perspective (2001–2010). Foodborne Pathog. Dis. 2014, 11, 425–432. [Google Scholar] [CrossRef]
  146. Bjarnsholt, T.; Jensen, P.Ø.; Rasmussen, T.B.; Christophersen, L.; Calum, H.; Hentzer, M.; Hougen, H.-P.; Rygaard, J.; Moser, C.; Eberl, L. Garlic blocks quorum sensing and promotes rapid clearing of pulmonary Pseudomonas aeruginosa infections. Microbiology 2005, 151, 3873–3880. [Google Scholar] [CrossRef] [Green Version]
  147. 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]
  148. Kazemian, H.; Ghafourian, S.; Heidari, H.; Amiri, P.; Yamchi, J.K.; Shavalipour, A.; Houri, H.; Maleki, A.; Sadeghifard, N. Antibacterial, anti-swarming and anti-biofilm formation activities of Chamaemelum nobile against Pseudomonas aeruginosa. Rev. Soc. Bras. Med. Trop. 2015, 48, 432–436. [Google Scholar] [CrossRef] [Green Version]
  149. Howell, A.B. Cranberry proanthocyanidins and the maintenance of urinary tract health. Crit. Rev. Food Sci. Nutr. 2002, 42, 273–278. [Google Scholar] [CrossRef]
  150. Burger, O.; Ofek, I.; Tabak, M.; Weiss, E.I.; Sharon, N.; Neeman, I. A high molecular mass constituent of cranberry juice inhibits Helicobacter pylori adhesion to human gastric mucus. FEMS Microbiol. Immunol. 2000, 29, 295–301. [Google Scholar] [CrossRef] [Green Version]
  151. Yamanaka, A.; Kimizuka, R.; Kato, T.; Okuda, K. Inhibitory effects of cranberry juice on attachment of oral streptococci and biofilm formation. Oral Microbiol. Immunol. 2004, 19, 150–154. [Google Scholar] [CrossRef] [PubMed]
  152. Steinberg, D.; Feldman, M.; Ofek, I.; Weiss, E.I. Cranberry high molecular weight constituents promote Streptococcus sobrinus desorption from artificial biofilm. Int. J. Antimicrob. Agents 2005, 25, 247–251. [Google Scholar] [CrossRef] [PubMed]
  153. Kim, S.-W.; Chang, I.-M.; Oh, K.-B. Inhibition of the bacterial surface protein anchoring transpeptidase sortase by medicinal plants. Biosci. Biotechnol. Biochem. 2002, 66, 2751–2754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Gong, L.; Zou, W.; Zheng, K.; Shi, B.; Liu, M. The Herba patriniae (Caprifoliaceae): A review on traditional uses, phytochemistry, pharmacology and quality control. J. Ethnopharmacol. 2021, 265, 113264. [Google Scholar] [CrossRef] [PubMed]
  155. Gerstmeier, J.; Seegers, J.; Witt, F.; Waltenberger, B.; Temml, V.; Rollinger, J.M.; Stuppner, H.; Koeberle, A.; Schuster, D.; Werz, O. Ginkgolic acid is a multi-target inhibitor of key enzymes in pro-inflammatory lipid mediator biosynthesis. Front. Pharmacol. 2019, 10, 797. [Google Scholar] [CrossRef]
  156. Lee, J.-H.; Kim, Y.-G.; Ryu, S.Y.; Cho, M.H.; Lee, J. Ginkgolic acids and Ginkgo biloba extract inhibit Escherichia coli O157: H7 and Staphylococcus aureus biofilm formation. Int. J. Food Microbiol. 2014, 174, 47–55. [Google Scholar] [CrossRef]
  157. Elekhnawy, E.; Negm, W.A.; El-Aasr, M.; Kamer, A.A.; Alqarni, M.; Batiha, G.E.-S.; Obaidullah, A.J.; Fawzy, H.M. Histological assessment, anti-quorum sensing, and anti-biofilm activities of Dioon spinulosum extract: In vitro and in vivo approach. Sci. Rep. 2022, 12, 180. [Google Scholar] [CrossRef]
  158. Obaid, R.F.; Kadhim Hindi, N.K.; Kadhum, S.A.; Jafaar Alwaeli, L.A.; Jalil, A.T. Antibacterial activity, anti-adherence and anti-biofilm activities of plants extracts against Aggregatibacter actinomycetemcomitans: An in vitro study in Hilla City, Iraq. Casp. J. Environ. Sci. 2022, 20, 367–372. [Google Scholar]
  159. Negm, W.A.; El-Aasr, M.; Attia, G.; Alqahtani, M.J.; Yassien, R.I.; Abo Kamer, A.; Elekhnawy, E. Promising antifungal activity of Encephalartos laurentianus de wild against Candida albicans clinical isolates: In vitro and in vivo effects on renal cortex of adult albino rats. J. Fungi 2022, 8, 426. [Google Scholar] [CrossRef]
  160. Olawuwo, O.S.; Famuyide, I.M.; McGaw, L.J. Antibacterial and antibiofilm activity of selected medicinal plant leaf extracts against pathogens implicated in poultry diseases. Front. Vet. Sci. 2022, 9, 820304. [Google Scholar] [CrossRef]
  161. Fathi, M.; Ghane, M.; Pishkar, L. Phytochemical composition, antibacterial, and antibiofilm activity of Malva sylvestris against human pathogenic bacteria. Jundishapur J. Nat. Pharm. Prod. 2022, 17, e114164. [Google Scholar] [CrossRef]
  162. Priyanto, J.A.; Prastya, M.E.; Sinarawadi, G.S.; Datu’salamah, W.; Avelina, T.Y.; Yanuar, A.I.A.; Azizah, E.; Tachrim, Z.P.; Mozef, T. The antibacterial and antibiofilm potential of Paederia foetida Linn. leaves extract. J. Appl. Pharm. Sci. 2022, 12, 117–124. [Google Scholar] [CrossRef]
  163. Panjaitan, C.C.; Widyarman, A.S.; Amtha, R.; Astoeti, T.E. Antimicrobial and antibiofilm activity of cinnamon (Cinnamomum burmanii) extract on periodontal pathogens—An in vitro study. Eur. J. Dent. 2022. Online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  164. Plescia, F.; Venturella, F.; Lauricella, M.; Catania, V.; Polito, G.; Schillaci, D.; Piccionello, A.P.; Giuseppe, D.; D’Anneo, A.; Raffa, D. Chemical composition, cytotoxic effects, antimicrobial and antibiofilm activity of Artemisia arborescens (Vaill.) L. growing wild in the province of Agrigento, Sicily, Italy. Plant Biosyst. 2022, 1–10. [Google Scholar] [CrossRef]
  165. Rhimi, W.; Mohammed, M.A.; Zarea, A.A.K.; Greco, G.; Tempesta, M.; Otranto, D.; Cafarchia, C. Antifungal, antioxidant and antibiofilm activities of essential oils of Cymbopogon spp. Antibiotics 2022, 11, 829. [Google Scholar] [CrossRef]
  166. Nazzaro, F.; Polito, F.; Amato, G.; Caputo, L.; Francolino, R.; D’Acierno, A.; Fratianni, F.; Candido, V.; Coppola, R.; De Feo, V. Chemical composition of essential oils of bulbs and aerial parts of two cultivars of Allium sativum and their antibiofilm activity against food and nosocomial pathogens. Antibiotics 2022, 11, 724. [Google Scholar] [CrossRef]
  167. Gamal El-Din, M.I.; Youssef, F.S.; Altyar, A.E.; Ashour, M.L. GC/MS Analyses of the essential oils obtained from different jatropha species, their discrimination using chemometric analysis and assessment of their antibacterial and anti-biofilm activities. Plants 2022, 11, 1268. [Google Scholar] [CrossRef]
  168. Djebili, S.; Taş, M.; Bouguerra, A.; Kucukaydin, S.; Ceylan, O.; Duru, M.E.; Barkat, M. Volatile compound profile and essential oil composition of three wild Algerian aromatic plants with their antioxidant and antibiofilm activities. J. Food Meas. Charact. 2022, 16, 987–999. [Google Scholar] [CrossRef]
  169. De Oliveira Galvão, F.; da Silva Dantas, F.G.; de Lima Santos, C.R.; Marchioro, S.B.; Cardoso, C.A.L.; Wender, H.; Sangalli, A.; de Almeida-Apolonio, A.A.; de Oliveira, K.M.P. Cochlospermum regium (Schrank) pilger leaf extract inhibit methicillin-resistant Staphylococcus aureus biofilm formation. J. Ethnopharmacol. 2020, 261, 113167. [Google Scholar] [CrossRef]
  170. Vijayakumar, K.; Ramanathan, T. Musa acuminata and its bioactive metabolite 5-Hydroxymethylfurfural mitigates quorum sensing (las and rhl) mediated biofilm and virulence production of nosocomial pathogen Pseudomonas aeruginosa in vitro. J. Ethnopharmacol. 2020, 246, 112242. [Google Scholar] [CrossRef]
  171. Tang, Y.; Bai, J.; Yang, Y.; Bai, X.; Bello-Onaghise, G.s.; Xu, Y.; Li, Y. Effect of syringopicroside extracted from syringa oblata lindl on the biofilm formation of Streptococcus suis. Molecules 2021, 26, 1295. [Google Scholar] [CrossRef] [PubMed]
  172. Bocquet, L.; Sahpaz, S.; Bonneau, N.; Beaufay, C.; Mahieux, S.; Samaillie, J.; Roumy, V.; Jacquin, J.; Bordage, S.; Hennebelle, T. Phenolic compounds from Humulus lupulus as natural antimicrobial products: New weapons in the fight against methicillin resistant Staphylococcus aureus, Leishmania mexicana and Trypanosoma brucei strains. Molecules 2019, 24, 1024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Attallah, N.G.; Al-Fakhrany, O.M.; Elekhnawy, E.; Hussein, I.A.; Shaldam, M.A.; Altwaijry, N.; Alqahtani, M.J.; Negm, W.A. Anti-biofilm and antibacterial activities of Cycas media R. Br secondary metabolites: In silico, in vitro, and in vivo approaches. Antibiotics 2022, 11, 993. [Google Scholar] [CrossRef] [PubMed]
  174. Viktorová, J.; Stupák, M.; Řehořová, K.; Dobiasová, S.; Hoang, L.; Hajšlová, J.; Van Thanh, T.; Van Tri, L.; Van Tuan, N.; Ruml, T. Lemon grass essential oil does not modulate cancer cells multidrug resistance by citral—Its dominant and strongly antimicrobial compound. Foods 2020, 9, 585. [Google Scholar] [CrossRef]
  175. Alibi, S.; Selma, W.B.; Ramos-Vivas, J.; Smach, M.A.; Touati, R.; Boukadida, J.; Navas, J.; Mansour, H.B. Anti-oxidant, antibacterial, anti-biofilm, and anti-quorum sensing activities of four essential oils against multidrug-resistant bacterial clinical isolates. Curr. Res. Transl. Med. 2020, 68, 59–66. [Google Scholar] [CrossRef]
  176. Umamaheswari, A.; SJ, G.F. Investigation on the biofilm eradication potential of selected medicinal plants against methicillin-resistant Staphylococcus aureus. Biotechnol. Rep. 2020, 28, e00523. [Google Scholar]
  177. Elamary, R.B.; Albarakaty, F.M.; Salem, W.M. Efficacy of Acacia nilotica aqueous extract in treating biofilm-forming and multidrug resistant uropathogens isolated from patients with UTI syndrome. Sci. Rep. 2020, 10, 11125. [Google Scholar] [CrossRef]
  178. Aumeeruddy-Elalfi, Z.; Ismaël, I.S.; Hosenally, M.; Zengin, G.; Mahomoodally, M.F. Essential oils from tropical medicinal herbs and food plants inhibit biofilm formation in vitro and are non-cytotoxic to human cells. 3 Biotech 2018, 8, 395. [Google Scholar] [CrossRef]
  179. Bernal-Mercado, A.T.; Vazquez-Armenta, F.J.; Tapia-Rodriguez, M.R.; Islas-Osuna, M.A.; Mata-Haro, V.; Gonzalez-Aguilar, G.A.; Lopez-Zavala, A.A.; Ayala-Zavala, J.F. Comparison of single and combined use of catechin, protocatechuic, and vanillic acids as antioxidant and antibacterial agents against uropathogenic Escherichia coli at planktonic and biofilm levels. Molecules 2018, 23, 2813. [Google Scholar] [CrossRef] [Green Version]
  180. Dias-Souza, M.V.; Dos Santos, R.M.; Cerávolo, I.P.; Cosenza, G.; Marçal, P.H.F. Euterpe oleracea pulp extract: Chemical analyses, antibiofilm activity against Staphylococcus aureus, cytotoxicity and interference on the activity of antimicrobial drugs. Microb. Pathog. 2018, 114, 29–35. [Google Scholar] [CrossRef]
  181. Dolatabadi, S.; Nesari, H.; Mahdavi-ourtakand, M. Microbial pathogenesis evaluating the anti-biofilm and antibacterial effects of Juglans regia L. extracts against clinical isolates of Pseudomonas aeruginosa. Microb. Pathog 2018, 118, 285–289. [Google Scholar] [CrossRef] [PubMed]
  182. Endo, E.H.; Costa, G.M.; Makimori, R.Y.; Ueda-Nakamura, T.; Nakamura, C.V.; Dias Filho, B.P. Anti-biofilm activity of Rosmarinus officinalis, Punica granatum and Tetradenia riparia against methicillin-resistant Staphylococcus aureus (MRSA) and synergic interaction with penicillin. J. Herb. Med. 2018, 14, 48–54. [Google Scholar] [CrossRef]
  183. Romero, C.M.; Vivacqua, C.G.; Abdulhamid, M.B.; Baigori, M.D.; Slanis, A.C.; Allori, M.C.G.d.; Tereschuk, M.L. Biofilm inhibition activity of traditional medicinal plants from Northwestern Argentina against native pathogen and environmental microorganisms. Rev. Soc. Bras. Med. Trop. 2016, 49, 703–712. [Google Scholar] [CrossRef] [PubMed]
  184. Cartagena, E.; Colom, O.Á.; Neske, A.; Valdez, J.C.; Bardón, A. Effects of plant lactones on the production of biofilm of Pseudomonas aeruginosa. Chem. Pharm. Bull. 2007, 55, 22–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Combarros-Fuertes, P.; Estevinho, L.M.; Dias, L.G.; Castro, J.M.; Tomás-Barberán, F.A.; Tornadijo, M.E.; Fresno-Baro, J.M. Bioactive components and antioxidant and antibacterial activities of different varieties of honey: A screening prior to clinical application. J. Agric. Food Chem. 2018, 67, 688–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Maddocks, S.E.; Jenkins, R.E. Honey: A sweet solution to the growing problem of antimicrobial resistance? Future Microbiol. 2013, 8, 1419–1429. [Google Scholar] [CrossRef]
  187. Cooper, R.; Jenkins, R. Are there feasible prospects for manuka honey as an alternative to conventional antimicrobials? Expert Rev. Anti-Infect. Ther. 2012, 10, 623–625. [Google Scholar] [CrossRef]
  188. Bouchelaghem, S.; Das, S.; Naorem, R.S.; Czuni, L.; Papp, G.; Kocsis, M. Evaluation of total phenolic and flavonoid contents, antibacterial and antibiofilm activities of Hungarian Propolis ethanolic extract against Staphylococcus aureus. Molecules 2022, 27, 574. [Google Scholar] [CrossRef]
  189. Alandejani, T.; Marsan, J.; Ferris, W.; Slinger, R.; Chan, F. Effectiveness of honey on Staphylococcus aureus and Pseudomonas aeruginosa biofilms. Otolaryngol. Head Neck Surg. 2009, 141, 114–118. [Google Scholar] [CrossRef]
  190. Qamar, M.U.; Saleem, S.; Toleman, M.A.; Saqalein, M.; Waseem, M.; Nisar, M.A.; Khurshid, M.; Taj, Z.; Jahan, S. In vitro and in vivo activity of Manuka honey against NDM-1-producing Klebsiella pneumoniae ST11. Future Microbiol. 2018, 13, 13–26. [Google Scholar] [CrossRef]
  191. Girma, A.; Seo, W.; She, R.C. Antibacterial activity of varying UMF-graded Manuka honeys. PLoS ONE 2019, 14, e0224495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Shah Pratibha, J.; Williamson Manita, T. Antibacterial activity of honey against ESBL producing Klebsiella pneumoniae from burn wound infections. Int. J. Curr. Pharm. Res. 2015, 7, 32–36. [Google Scholar]
  193. Idris, A.R.; Afegbua, S.L. Single and joint antibacterial activity of aqueous garlic extract and Manuka honey on extended-spectrum beta-lactamase-producing Escherichia coli. Trans. R. Soc. Trop. Med. Hyg. 2017, 111, 472–478. [Google Scholar] [CrossRef]
  194. Hillitt, K.; Jenkins, R.; Spiller, O.B.; Beeton, M.L. Antimicrobial activity of Manuka honey against antibiotic-resistant strains of the cell wall-free bacteria Ureaplasma parvum and Ureaplasma urealyticum. Lett. Appl. Microbiol. 2017, 64, 198–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Fadl, A.E.-W. Antibacterial and antibiofilm effects of bee venom from (Apis mellifera) on multidrug-resistant bacteria (MDRB). Al-Azhar J. Pharm. Sci. 2018, 58, 60–80. [Google Scholar] [CrossRef]
  196. Jeevanandam, J.; Aing, Y.S.; Chan, Y.S.; Pan, S.; Danquah, M.K. Nanoformulation and application of phytochemicals as antimicrobial agents. In Antimicrobial Nanoarchitectonics; Elsevier: Amsterdam, The Netherlands, 2017; pp. 61–82. [Google Scholar]
  197. Cartaxo, A.L.P. Nanoparticles types and properties–understanding these promising devices in the biomedical area. Biol. Mater. Sci. 2015, 1–8. Available online: https://www.semanticscholar.org/paper/Nanoparticles-types-and-properties-%E2%80%93-understanding-Cartaxo/9570b62e051cf8cff1398915cf14eafcff21fbdd?sort=is-influential (accessed on 11 October 2022).
  198. Wang, L.; Hu, C.; Shao, L. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomed. 2017, 12, 1227. [Google Scholar] [CrossRef] [Green Version]
  199. Zhang, K.; Li, X.; Yu, C.; Wang, Y. Promising Therapeutic Strategies against Microbial Biofilm Challenges. Front. Cell. Infect. Microbiol. 2020, 10, 359. [Google Scholar] [CrossRef]
  200. Ahmad, N.; Bhatnagar, S.; Ali, S.S.; Dutta, R. Phytofabrication of bioinduced silver nanoparticles for biomedical applications. Int. J. Nanomed. 2015, 10, 7019. [Google Scholar]
  201. Swidan, N.S.; Hashem, Y.A.; Elkhatib, W.F.; Yassien, M.A. Antibiofilm activity of green synthesized silver nanoparticles against biofilm associated enterococcal urinary pathogens. Sci. Rep. 2022, 12, 3869. [Google Scholar] [CrossRef]
  202. Muthulakshmi, L.; Suganya, K.; Murugan, M.; Annaraj, J.; Duraipandiyan, V.; Al Farraj, D.A.; Elshikh, M.S.; Juliet, A.; Pasupuleti, M.; Arockiaraj, J. Antibiofilm efficacy of novel biogenic silver nanoparticles from Terminalia catappa against food-borne Listeria monocytogenes ATCC 15313 and mechanisms investigation in-vivo and in-vitro. J. King Saud Univ. Sci. 2022, 34, 102083. [Google Scholar] [CrossRef]
  203. Salem, S.S.; Badawy, M.S.E.; Al-Askar, A.A.; Arishi, A.A.; Elkady, F.M.; Hashem, A.H. Green biosynthesis of selenium nanoparticles using orange peel waste: Characterization, antibacterial and antibiofilm activities against multidrug-resistant bacteria. Life 2022, 12, 893. [Google Scholar] [CrossRef] [PubMed]
  204. Rajivgandhi, G.N.; Maruthupandy, M.; Li, J.-L.; Dong, L.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Alanzi, K.F.; Li, W.-J. Photocatalytic reduction and anti-bacterial activity of biosynthesized silver nanoparticles against multi drug resistant Staphylococcus saprophyticus BDUMS 5 (MN310601). Mater. Sci. Eng. C. 2020, 114, 111024. [Google Scholar] [CrossRef] [PubMed]
  205. Mohanta, Y.K.; Biswas, K.; Jena, S.K.; Hashem, A.; Abd_Allah, E.F.; Mohanta, T.K. Anti-biofilm and antibacterial activities of silver nanoparticles synthesized by the reducing activity of phytoconstituents present in the Indian medicinal plants. Front. Microbiol. 2020, 11, 1143. [Google Scholar] [CrossRef] [PubMed]
  206. Mariadoss, A.V.A.; Ramachandran, V.; Shalini, V.; Agilan, B.; Franklin, J.H.; Sanjay, K.; Alaa, Y.G.; Tawfiq, M.A.-A.; Ernest, D. Green synthesis, characterization and antibacterial activity of silver nanoparticles by Malus domestica and its cytotoxic effect on (MCF-7) cell line. Microb. Pathog. 2019, 135, 103609. [Google Scholar] [CrossRef]
  207. Shah, S.; Gaikwad, S.; Nagar, S.; Kulshrestha, S.; Vaidya, V.; Nawani, N.; Pawar, S. Biofilm inhibition and anti-quorum sensing activity of phytosynthesized silver nanoparticles against the nosocomial pathogen Pseudomonas aeruginosa. Biofouling 2019, 35, 34–49. [Google Scholar] [CrossRef] [Green Version]
  208. Tian, X.; Wang, P.; Li, T.; Huang, X.; Guo, W.; Yang, Y.; Yan, M.; Zhang, H.; Cai, D.; Jia, X. Self-assembled natural phytochemicals for synergistically antibacterial application from the enlightenment of traditional Chinese medicine combination. Acta Pharm. Sin. B. 2020, 10, 1784–1795. [Google Scholar] [CrossRef]
  209. Huang, X.; Wang, P.; Li, T.; Tian, X.; Guo, W.; Xu, B.; Huang, G.; Cai, D.; Zhou, F.; Zhang, H. Self-assemblies based on traditional medicine berberine and cinnamic acid for adhesion-induced inhibition multidrug-resistant Staphylococcus aureus. ACS Appl. Mater. Interfaces 2019, 12, 227–237. [Google Scholar] [CrossRef]
  210. Cherian, T.; Ali, K.; Saquib, Q.; Faisal, M.; Wahab, R.; Musarrat, J. Cymbopogon citratus functionalized green synthesis of CuO-nanoparticles: Novel prospects as antibacterial and antibiofilm agents. Biomolecules 2020, 10, 169. [Google Scholar] [CrossRef] [Green Version]
  211. Lotha, R.; Shamprasad, B.R.; Sundaramoorthy, N.S.; Nagarajan, S.; Sivasubramanian, A. Biogenic phytochemicals (cassinopin and isoquercetin) capped copper nanoparticles (ISQ/CAS@ CuNPs) inhibits MRSA biofilms. Microb. Pathog. 2019, 132, 178–187. [Google Scholar] [CrossRef]
  212. Cherian, T.; Ali, K.; Fatima, S.; Saquib, Q.; Ansari, S.M.; Alwathnani, H.A.; Al-Khedhairy, A.A.; Al-Shaeri, M.; Musarrat, J. Myristica fragrans bio-active ester functionalized ZnO nanoparticles exhibit antibacterial and antibiofilm activities in clinical isolates. J. Microbiol. Methods 2019, 166, 105716. [Google Scholar] [CrossRef] [PubMed]
  213. Scriboni, A.B.; Couto, V.M.; Ribeiro, L.N.D.M.; Freires, I.A.; Groppo, F.C.; De Paula, E.; Franz-Montan, M.; Cogo-Müller, K. Fusogenic liposomes increase the antimicrobial activity of vancomycin against Staphylococcus aureus biofilm. Front. Pharmacol. 2019, 10, 1401. [Google Scholar] [CrossRef] [PubMed]
  214. Zahra, M.-J.; Hamed, H.; Mohammad, R.-Y.; Nosratollah, Z.; Akbarzadeh, A.; Morteza, M. Evaluation and study of antimicrobial activity of nanoliposomal meropenem against Pseudomonas aeruginosa isolates. Artif. Cells Nanomed. Biotechnol. 2017, 45, 975–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Paliwal, R.; Paliwal, S.R.; Kenwat, R.; Kurmi, B.D.; Sahu, M.K. Solid lipid nanoparticles: A review on recent perspectives and patents. Expert Opin. Ther. Pat. 2020, 30, 179–194. [Google Scholar] [CrossRef]
  216. Singh, B.; Vuddanda, P.R.; Vijayakumar, M.; Kumar, V.; Saxena, P.S.; Singh, S. Cefuroxime axetil loaded solid lipid nanoparticles for enhanced activity against S. aureus biofilm. Colloids Surf. B. 2014, 121, 92–98. [Google Scholar] [CrossRef]
  217. Nafee, N.; Husari, A.; Maurer, C.K.; Lu, C.; de Rossi, C.; Steinbach, A.; Hartmann, R.W.; Lehr, C.-M.; Schneider, M. Antibiotic-free nanotherapeutics: Ultra-small, mucus-penetrating solid lipid nanoparticles enhance the pulmonary delivery and anti-virulence efficacy of novel quorum sensing inhibitors. J. Control. Release 2014, 192, 131–140. [Google Scholar] [CrossRef]
  218. Gómez-Gómez, B.; Arregui, L.; Serrano, S.; Santos, A.; Pérez-Corona, T.; Madrid, Y. Selenium and tellurium-based nanoparticles as interfering factors in quorum sensing-regulated processes: Violacein production and bacterial biofilm formation. Metallomics 2019, 11, 1104–1114. [Google Scholar] [CrossRef]
  219. Nguyen, H.T.; Hensel, A.; Goycoolea, F.M. Chitosan/cyclodextrin surface-adsorbed naringenin-loaded nanocapsules enhance bacterial quorum quenching and anti-biofilm activities. Colloids Surf. B. 2022, 211, 112281. [Google Scholar] [CrossRef]
  220. Nava-Ortiz, C.A.; Burillo, G.; Concheiro, A.; Bucio, E.; Matthijs, N.; Nelis, H.; Coenye, T.; Alvarez-Lorenzo, C. Cyclodextrin-functionalized biomaterials loaded with miconazole prevent Candida albicans biofilm formation in vitro. Acta Biomater. 2010, 6, 1398–1404. [Google Scholar] [CrossRef]
  221. Gharbi, A.; Humblot, V.; Turpin, F.; Pradier, C.-M.; Imbert, C.; Berjeaud, J.-M. Elaboration of antibiofilm surfaces functionalized with antifungal-cyclodextrin inclusion complexes. FEMS Microbiol. Immunol. 2012, 65, 257–269. [Google Scholar] [CrossRef]
  222. Huang, Y.; Bai, L.; Yang, Y.; Yin, Z.; Guo, B. Biodegradable gelatin/silver nanoparticle composite cryogel with excellent antibacterial and antibiofilm activity and hemostasis for Pseudomonas aeruginosa-infected burn wound healing. J. Colloid Interface Sci. 2022, 608, 2278–2289. [Google Scholar] [CrossRef] [PubMed]
  223. Haidari, H.; Bright, R.; Garg, S.; Vasilev, K.; Cowin, A.J.; Kopecki, Z. Eradication of mature bacterial biofilms with concurrent improvement in chronic wound healing using silver nanoparticle hydrogel treatment. Biomedicines 2021, 9, 1182. [Google Scholar] [CrossRef] [PubMed]
  224. Hemmingsen, L.M.; Giordani, B.; Pettersen, A.K.; Vitali, B.; Basnet, P.; Škalko-Basnet, N. Liposomes-in-chitosan hydrogel boosts potential of chlorhexidine in biofilm eradication in vitro. Carbohydr. Polym. 2021, 262, 117939. [Google Scholar] [CrossRef]
  225. Li, X.; Fu, Y.-n.; Huang, L.; Liu, F.; Moriarty, T.F.; Tao, L.; Wei, Y.; Wang, X. Combating biofilms by a self-adapting drug loading hydrogel. ACS Appl. Bio Mater. 2021, 4, 6219–6226. [Google Scholar] [CrossRef] [PubMed]
  226. Ouyang, J.; Bu, Q.; Tao, N.; Chen, M.; Liu, H.; Zhou, J.; Liu, J.; Deng, B.; Kong, N.; Zhang, X. A facile and general method for synthesis of antibiotic-free protein-based hydrogel: Wound dressing for the eradication of drug-resistant bacteria and biofilms. Bioact. Mater. 2022, 18, 446–458. [Google Scholar] [CrossRef]
  227. Chen, H.; Cheng, J.; Cai, X.; Han, J.; Chen, X.; You, L.; Xiong, C.; Wang, S. pH-Switchable antimicrobial supramolecular hydrogels for synergistically eliminating biofilm and promoting wound healing. ACS Appl. Mater. Interfaces 2022, 14, 18120–18132. [Google Scholar] [CrossRef]
  228. Alfuraydi, R.T.; Alminderej, F.M.; Mohamed, N.A. Evaluation of antimicrobial and anti-biofilm formation activities of novel poly (vinyl alcohol) hydrogels reinforced with crosslinked chitosan and silver nano-particles. Polymers 2022, 14, 1619. [Google Scholar] [CrossRef]
  229. Wroe, J.A.; Johnson, C.T.; García, A.J. Bacteriophage delivering hydrogels reduce biofilm formation in vitro and infection in vivo. J. Biomed. Mater. Res. A 2020, 108, 39–49. [Google Scholar] [CrossRef]
  230. Barros, J.A.R.; de Melo, L.D.R.; da Silva, R.A.R.; Ferraz, M.P.; de Rodrigues Azeredo, J.C.V.; de Carvalho Pinheiro, V.M.; Colaço, B.J.A.; Fernandes, M.H.R.; de Sousa Gomes, P.; Monteiro, F.J. Encapsulated bacteriophages in alginate-nanohydroxyapatite hydrogel as a novel delivery system to prevent orthopedic implant-associated infections. Nanomed. Nanotechnol. Biol. Med. 2020, 24, 102145. [Google Scholar] [CrossRef] [Green Version]
  231. Tarawneh, O.; Abu Mahfouz, H.; Hamadneh, L.; Deeb, A.A.; Al-Sheikh, I.; Alwahsh, W.; Fadhil Abed, A. Assessment of persistent antimicrobial and anti-biofilm activity of p-HEMA hydrogel loaded with rifampicin and cefixime. Sci. Rep. 2022, 12, 3900. [Google Scholar] [CrossRef]
  232. Anjum, A.; Sim, C.-H.; Ng, S.-F. Hydrogels containing antibiofilm and antimicrobial agents beneficial for biofilm-associated wound infection: Formulation characterizations and In vitro study. AAPS PharmSciTech 2018, 19, 1219–1230. [Google Scholar] [CrossRef] [PubMed]
  233. Leclercq, L.; Tessier, J.; Douyere, G.; Nardello-Rataj, V.; Schmitzer, A.R. Phytochemical-and cyclodextrin-based pickering emulsions: Natural potentiators of antibacterial, antifungal, and antibiofilm activity. Langmuir 2020, 36, 4317–4323. [Google Scholar] [CrossRef] [PubMed]
  234. Tiwari, G.; Tiwari, R.; Rai, A.K. Cyclodextrins in delivery systems: Applications. J. Pharm. Bioallied Sci. 2010, 2, 72. [Google Scholar] [CrossRef]
  235. Agrahari, A.K.; Singh, A.K.; Singh, A.S.; Singh, M.; Maji, P.; Yadav, S.; Rajkhowa, S.; Prakash, P.; Tiwari, V.K. Click inspired synthesis of p-tert-butyl calix [4] arene tethered benzotriazolyl dendrimers and their evaluation as anti-bacterial and anti-biofilm agents. New J. Chem. 2020, 44, 19300–19313. [Google Scholar] [CrossRef]
  236. Barik, S.K.; Singh, B.N. Nanoemulsion-loaded hydrogel coatings for inhibition of bacterial virulence and biofilm formation on solid surfaces. Sci. Rep. 2019, 9, 6520. [Google Scholar]
  237. Liang, Y.; He, J.; Guo, B. Functional hydrogels as wound dressing to enhance wound healing. ACS Nano 2021, 15, 12687–12722. [Google Scholar] [CrossRef]
  238. Dong, D.; Thomas, N.; Thierry, B.; Vreugde, S.; Clive, A.; Prestidge, C.A.; Wormald, P.-J. Distribution and inhibition of liposomes on Staphylococcus aureus and Pseudomonas aeruginosa biofilm. PLoS ONE 2015, 10, e0131806. [Google Scholar] [CrossRef] [Green Version]
  239. Wang, Y. Liposome as a delivery system for the treatment of biofilm-mediated infections. J. Appl. Microbiol. 2021, 131, 2626–2639. [Google Scholar] [CrossRef]
  240. Wang, C.; Chen, P.; Qiao, Y.; Kang, Y.; Yan, C.; Yu, Z.; Wang, J.; He, X.; Wu, H. pH responsive superporogen combined with PDT based on poly Ce6 ionic liquid grafted on SiO2 for combating MRSA biofilm infection. Theranostics 2020, 10, 4795. [Google Scholar] [CrossRef]
  241. Zhao, Z.; Ding, C.; Wang, Y.; Tan, H.; Li, J. pH-Responsive polymeric nanocarriers for efficient killing of cariogenic bacteria in biofilms. Biomater. Sci. 2019, 7, 1643–1651. [Google Scholar] [CrossRef]
  242. Singh, P.; Pandit, S.; Beshay, M.; Mokkapati, V.; Garnaes, J.; Olsson, M.E.; Sultan, A.; Mackevica, A.; Mateiu, R.V.; Lütken, H. Anti-biofilm effects of gold and silver nanoparticles synthesized by the Rhodiola rosea rhizome extracts. Artif. Cells Nanomed. Biotechnol. 2018, 46, S886–S899. [Google Scholar] [CrossRef]
  243. Liu, Y.; Shi, L.; Su, L.; van der Mei, H.C.; Jutte, P.C.; Ren, Y.; Busscher, H.J. Nanotechnology-based antimicrobials and delivery systems for biofilm-infection control. Chem. Soc. Rev. 2019, 48, 428–446. [Google Scholar] [CrossRef] [PubMed]
  244. Szerencsés, B.; Igaz, N.; Tóbiás, Á.; Prucsi, Z.; Rónavári, A.; Bélteky, P.; Madarász, D.; Papp, C.; Makra, I.; Vágvölgyi, C. Size-dependent activity of silver nanoparticles on the morphological switch and biofilm formation of opportunistic pathogenic yeasts. BMC Microbiol. 2020, 20, 176. [Google Scholar] [CrossRef] [PubMed]
  245. Pham, P.; Oliver, S.; Wong, E.H.; Boyer, C. Effect of hydrophilic groups on the bioactivity of antimicrobial polymers. Polym. Chem. 2021, 12, 5689–5703. [Google Scholar] [CrossRef]
  246. Olmo, J.A.-D.; Ruiz-Rubio, L.; Pérez-Alvarez, L.; Sáez-Martínez, V.; Vilas-Vilela, J.L. Antibacterial coatings for improving the performance of biomaterials. Coatings 2020, 10, 139. [Google Scholar] [CrossRef] [Green Version]
  247. Yu, Q.; Li, J.; Zhang, Y.; Wang, Y.; Liu, L.; Li, M. Inhibition of gold nanoparticles (AuNPs) on pathogenic biofilm formation and invasion to host cells. Sci. Rep. 2016, 6, 26667. [Google Scholar] [CrossRef]
  248. Giri, K.; Yepes, L.R.; Duncan, B.; Parameswaran, P.K.; Yan, B.; Jiang, Y.; Bilska, M.; Moyano, D.F.; Thompson, M.A.; Rotello, V.M. Targeting bacterial biofilms via surface engineering of gold nanoparticles. RSC Adv. 2015, 5, 105551–105559. [Google Scholar] [CrossRef] [Green Version]
  249. Sathyanarayanan, M.B.; Balachandranath, R.; Genji Srinivasulu, Y.; Kannaiyan, S.K.; Subbiahdoss, G. The effect of gold and iron-oxide nanoparticles on biofilm-forming pathogens. Int. Sch. Res. Not. 2013, 2013, 272086. [Google Scholar] [CrossRef] [Green Version]
  250. Van de Walle, A.; Perez, J.E.; Abou-Hassan, A.; Hémadi, M.; Luciani, N.; Wilhelm, C. Magnetic nanoparticles in regenerative medicine: What of their fate and impact in stem cells? Mater. Today Nano 2020, 11, 100084. [Google Scholar] [CrossRef]
  251. Belkahla, H.; Mazarío, E.; Sangnier, A.P.; Lomas, J.S.; Gharbi, T.; Ammar, S.; Micheau, O.; Wilhelm, C.; Hémadi, M. TRAIL acts synergistically with iron oxide nanocluster-mediated magneto-and photothermia. Theranostics 2019, 9, 5924. [Google Scholar] [CrossRef]
  252. Di Corato, R.; Béalle, G.; Kolosnjaj-Tabi, J.; Espinosa, A.; Clement, O.; Silva, A.K.; Menager, C.; Wilhelm, C. Combining magnetic hyperthermia and photodynamic therapy for tumor ablation with photoresponsive magnetic liposomes. ACS Nano 2015, 9, 2904–2916. [Google Scholar] [CrossRef] [PubMed]
  253. Hasanzadeh, M.; Shadjou, N.; de la Guardia, M. Iron and iron-oxide magnetic nanoparticles as signal-amplification elements in electrochemical biosensing. TrAC Trends Anal. Chem. 2015, 72, 1–9. [Google Scholar] [CrossRef]
  254. Kim, S.H.; Kwak, S.-Y.; Sohn, B.-H.; Park, T.H. Design of TiO2 nanoparticle self-assembled aromatic polyamide thin-film-composite (TFC) membrane as an approach to solve biofouling problem. J. Membr. Sci. 2003, 211, 157–165. [Google Scholar] [CrossRef]
  255. Scioli Montoto, S.; Muraca, G.; Ruiz, M.E. Solid lipid nanoparticles for drug delivery: Pharmacological and biopharmaceutical aspects. Front. Mol. Biosci. 2020, 7, 587997. [Google Scholar] [CrossRef]
  256. Liu, Y.; Naha, P.C.; Hwang, G.; Kim, D.; Huang, Y.; Simon-Soro, A.; Jung, H.-I.; Ren, Z.; Li, Y.; Gubara, S. Topical ferumoxytol nanoparticles disrupt biofilms and prevent tooth decay in vivo via intrinsic catalytic activity. Nat. Commun. 2018, 9, 2920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Qayyum, S.; Oves, M.; Khan, A.U. Obliteration of bacterial growth and biofilm through ROS generation by facilely synthesized green silver nanoparticles. PLoS ONE 2017, 12, e0181363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Siddique, M.H.; Aslam, B.; Imran, M.; Ashraf, A.; Nadeem, H.; Hayat, S.; Khurshid, M.; Afzal, M.; Malik, I.R.; Shahzad, M. Effect of silver nanoparticles on biofilm formation and EPS production of multidrug-resistant Klebsiella pneumoniae. Biomed Res. Int. 2020, 2020, 6398165. [Google Scholar] [CrossRef] [Green Version]
  259. Miller, K.P.; Wang, L.; Chen, Y.-P.; Pellechia, P.J.; Benicewicz, B.C.; Decho, A.W. Engineering nanoparticles to silence bacterial communication. Front. Microbiol. 2015, 6, 189. [Google Scholar] [CrossRef] [Green Version]
  260. Mangzira Kemung, H.; Tan, L.T.-H.; Chan, K.-G.; Ser, H.-L.; Law, J.W.-F.; Lee, L.-H.; Goh, B.-H. Streptomyces sp. strain MUSC 125 from mangrove soil in Malaysia with anti-MRSA, anti-biofilm and antioxidant activities. Molecules 2020, 25, 3545. [Google Scholar] [CrossRef] [PubMed]
  261. Younis, K.M.; Usup, G.; Ahmad, A. Secondary metabolites produced by marine streptomyces as antibiofilm and quorum-sensing inhibitor of uropathogen Proteus mirabilis. Environ. Sci. Pollut. Res. 2016, 23, 4756–4767. [Google Scholar] [CrossRef]
  262. Peach, K.C.; Cheng, A.T.; Oliver, A.G.; Yildiz, F.H.; Linington, R.G. Discovery and biological characterization of the auromomycin chromophore as an inhibitor of biofilm formation in Vibrio cholerae. Chembiochem 2013, 14, 2209–2215. [Google Scholar] [CrossRef] [PubMed]
  263. Manefield, M.; de Nys, R.; Naresh, K.; Roger, R.; Givskov, M.; Peter, S.; Kjelleberg, S. Evidence that halogenated furanones from Delisea pulchra inhibit acylated homoserine lactone (AHL)-mediated gene expression by displacing the AHL signal from its receptor protein. Microbiology 1999, 145, 283–291. [Google Scholar] [CrossRef] [Green Version]
  264. Ren, D.; Bedzyk, L.A.; Setlow, P.; England, D.F.; Kjelleberg, S.; Thomas, S.M.; Ye, R.W.; Wood, T.K. Differential gene expression to investigate the effect of (5Z)-4-bromo-5-(bromomethylene)-3-butyl-2 (5H)-furanone on Bacillus subtilis. Appl. Environ. Microbiol. 2004, 70, 4941–4949. [Google Scholar] [CrossRef] [Green Version]
  265. Ren, D.; Sims, J.; Wood, T. Inhibition of biofilm formation and swarming of Bacillus subtilis by (5Z)-4-bromo-5-(bromomethylene)-3-butyl-2 (5H)-furanone. Lett. Appl. Microbiol. 2002, 34, 293–299. [Google Scholar] [CrossRef] [Green Version]
  266. Manefield, M.; Rasmussen, T.B.; Henzter, M.; Andersen, J.B.; Steinberg, P.; Kjelleberg, S.; Givskov, M. Halogenated furanones inhibit quorum sensing through accelerated LuxR turnover. Microbiology 2002, 148, 1119–1127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Sahreen, S.; Mukhtar, H.; Imre, K.; Morar, A.; Herman, V.; Sharif, S. Exploring the function of quorum sensing regulated biofilms in biological wastewater treatment: A review. Int. J. Mol. Sci. 2022, 23, 9751. [Google Scholar] [CrossRef] [PubMed]
  268. Pereira, U.A.; Barbosa, L.C.; Maltha, C.R.; Demuner, A.J.; Masood, M.A.; Pimenta, A.L. γ-Alkylidene-γ-lactones and isobutylpyrrol-2 (5H)-ones analogues to rubrolides as inhibitors of biofilm formation by gram-positive and gram-negative bacteria. Bioorg. Med. Chem. Lett. 2014, 24, 1052–1056. [Google Scholar] [CrossRef] [PubMed]
  269. Tello, E.; Castellanos, L.; Arevalo-Ferro, C.; Rodríguez, J.; Jiménez, C.; Duque, C. Absolute stereochemistry of antifouling cembranoid epimers at C-8 from the Caribbean octocoral Pseudoplexaura flagellosa. Revised structures of plexaurolones. Tetrahedron 2011, 67, 9112–9121. [Google Scholar] [CrossRef]
  270. Barka, E.A.; Vatsa, P.; Sanchez, L.; Gaveau-Vaillant, N.; Jacquard, C.; Klenk, H.-P.; Clément, C.; Ouhdouch, Y.; van Wezel, G.P. Taxonomy, physiology, and natural products of Actinobacteria. Microbiol. Mol. Biol. Rev. 2016, 80, 1–43. [Google Scholar] [CrossRef] [Green Version]
  271. Song, Y.; Cai, Z.H.; Lao, Y.M.; Jin, H.; Ying, K.Z.; Lin, G.H.; Zhou, J. Antibiofilm activity substances derived from coral symbiotic bacterial extract inhibit biofouling by the model strain Pseudomonas aeruginosa PAO 1. Microb. Biotechnol. 2018, 11, 1090–1105. [Google Scholar] [CrossRef] [Green Version]
  272. Chen, X.; Chen, J.; Yan, Y.; Chen, S.; Xu, X.; Zhang, H.; Wang, H. Quorum sensing inhibitors from marine bacteria Oceanobacillus sp. XC22919. Nat. Prod. Res. 2019, 33, 1819–1823. [Google Scholar] [CrossRef] [PubMed]
  273. Goel, N.; Fatima, S.W.; Kumar, S.; Sinha, R.; Khare, S.K. Antimicrobial resistance in biofilms: Exploring marine actinobacteria as a potential source of antibiotics and biofilm inhibitors. Biotechnol. Rep. 2021, 30, e00613. [Google Scholar] [CrossRef] [PubMed]
  274. Marappa, N.; Dharumadurai, D.; Nooruddin, T.; Abdulkader, A.M. Morphological, molecular characterization and biofilm inhibition effect of endophytic Frankia sp. from root nodules of Actinorhizal plant Casuarina sp. S. Afr. J. Bot. 2020, 134, 72–83. [Google Scholar] [CrossRef]
  275. Singh, R.; Dubey, A.K. Isolation and characterization of a new endophytic actinobacterium Streptomyces californicus strain ADR1 as a promising source of anti-bacterial, anti-biofilm and antioxidant metabolites. Microorganisms 2020, 8, 929. [Google Scholar] [CrossRef] [PubMed]
  276. Kamarudheen, N.; Naushad, T.; Rao, K.V.B. Biosynthesis, characterization and antagonistic applications of extracellular melanin pigment from marine Nocardiopsis Sps. Ind. J. Pharm. Educ. Res 2019, 53, 112–120. [Google Scholar] [CrossRef] [Green Version]
  277. Ramalingam, V.; Mahamuni, D.; Rajaram, R. In vitro and in silico approaches of antibiofilm activity of 1-hydroxy-1-norresistomycin against human clinical pathogens. Microb. Pathog. 2019, 132, 343–354. [Google Scholar] [CrossRef]
  278. Rajivgandhi, G.; Vijayan, R.; Maruthupandy, M.; Vaseeharan, B.; Manoharan, N. Antibiofilm effect of Nocardiopsis sp. GRG 1 (KT235640) compound against biofilm forming Gram negative bacteria on UTIs. Microb. Pathog. 2018, 118, 190–198. [Google Scholar] [CrossRef]
  279. Miao, L.; Xu, J.; Yao, Z.; Jiang, Y.; Zhou, H.; Jiang, W.; Dong, K. The anti-quorum sensing activity and bioactive substance of a marine derived Streptomyces. Biotechnol. Biotechnol. Equip. 2017, 31, 1007–1015. [Google Scholar] [CrossRef] [Green Version]
  280. Chang, H.; Zhou, J.; Zhu, X.; Yu, S.; Chen, L.; Jin, H.; Cai, Z. Strain identification and quorum sensing inhibition characterization of marine-derived Rhizobium sp. NAO1. R. Soc. Open Sci. 2017, 4, 170025. [Google Scholar] [CrossRef] [Green Version]
  281. Stumpp, N.; Premnath, P.; Schmidt, T.; Ammermann, J.; Draeger, G.; Reck, M.; Jansen, R.; Stiesch, M.; Wagner-Döbler, I.; Kirschning, A. Synthesis of new carolacton derivatives and their activity against biofilms of oral bacteria. Org. Biomol. Chem. 2015, 13, 5765–5774. [Google Scholar] [CrossRef] [Green Version]
  282. Shin, D.-S.; Eom, Y.-B. Antimicrobial and antibiofilm activities of Clostridium butyricum supernatant against Acinetobacter baumannii. Arch. Microbiol. 2020, 202, 1059–1068. [Google Scholar] [CrossRef] [PubMed]
  283. Shokri, D.; Khorasgani, M.R.; Mohkam, M.; Fatemi, S.M.; Ghasemi, Y.; Taheri-Kafrani, A. The inhibition effect of lactobacilli against growth and biofilm formation of Pseudomonas aeruginosa. Probiotics Antimicrob. 2018, 10, 34–42. [Google Scholar] [CrossRef]
  284. Song, H.; Zhang, J.; Qu, J.; Liu, J.; Yin, P.; Zhang, G.; Shang, D. Lactobacillus rhamnosus GG microcapsules inhibit Escherichia coli biofilm formation in coculture. Biotechnol. Lett. 2019, 41, 1007–1014. [Google Scholar] [CrossRef] [PubMed]
  285. Roscetto, E.; Masi, M.; Esposito, M.; Di Lecce, R.; Delicato, A.; Maddau, L.; Calabrò, V.; Evidente, A.; Catania, M.R. Anti-biofilm activity of the fungal phytotoxin sphaeropsidin A against clinical isolates of antibiotic-resistant bacteria. Toxins 2020, 12, 444. [Google Scholar] [CrossRef]
  286. Petersen, L.-E.; Marner, M.; Labes, A.; Tasdemir, D. Rapid metabolome and bioactivity profiling of fungi associated with the leaf and rhizosphere of the Baltic seagrass Zostera marina. Mar. Drugs 2019, 17, 419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Narmani, A.; Teponno, R.B.; Helaly, S.E.; Arzanlou, M.; Stadler, M. Cytotoxic, anti-biofilm and antimicrobial polyketides from the plant associated fungus Chaetosphaeronema achilleae. Fitoterapia 2019, 139, 104390. [Google Scholar] [CrossRef]
  288. Arora, P.; Wani, Z.A.; Nalli, Y.; Ali, A.; Riyaz-Ul-Hassan, S. Antimicrobial potential of thiodiketopiperazine derivatives produced by Phoma sp., an endophyte of Glycyrrhiza glabra Linn. Microb. Ecol. 2016, 72, 802–812. [Google Scholar] [CrossRef]
  289. Rashmi, M.; Meena, H.; Meena, C.; Kushveer, J.; Busi, S.; Murali, A.; Sarma, V. Anti-quorum sensing and antibiofilm potential of Alternaria alternata, a foliar endophyte of Carica papaya, evidenced by QS assays and in-silico analysis. Fungal Biol. 2018, 122, 998–1012. [Google Scholar] [CrossRef]
  290. Zhang, M.; Wang, M.; Zhu, X.; Yu, W.; Gong, Q. Equisetin as potential quorum sensing inhibitor of Pseudomonas aeruginosa. Biotechnol. Lett. 2018, 40, 865–870. [Google Scholar] [CrossRef]
Figure 1. The development phases of biofilms.
Figure 1. The development phases of biofilms.
Life 12 01618 g001
Figure 2. Chemical structures of some signaling molecules.
Figure 2. Chemical structures of some signaling molecules.
Life 12 01618 g002
Figure 3. The life cycle of biofilm formation provides different intervention points for biofilm inhibition and eradication.
Figure 3. The life cycle of biofilm formation provides different intervention points for biofilm inhibition and eradication.
Life 12 01618 g003
Figure 4. Compounds that can inhibit biofilm formation.
Figure 4. Compounds that can inhibit biofilm formation.
Life 12 01618 g004
Figure 5. Schematic illustration of the mode of action of the fusogenic liposome.
Figure 5. Schematic illustration of the mode of action of the fusogenic liposome.
Life 12 01618 g005
Figure 6. Chemical structures of some brominated alkylidene lactams.
Figure 6. Chemical structures of some brominated alkylidene lactams.
Life 12 01618 g006
Table 1. List of a few identified polysaccharides in microbial biofilms.
Table 1. List of a few identified polysaccharides in microbial biofilms.
PolysaccharidesMicrobesRoleRef.
PGA
(Poly-β-1,6-N-acetyl-D-glucosamine)
Actinobacillus actinomycetemcomitansIntercellular adhesion
Cellular detachment
Dispersion
[28]
Colanic acidE. coliAntidesiccative[29]
Galactopyranosyl-glycerol-phosphateBacillus licheniformisAntibiofilm[30]
Alginate, Pel, PslP. aeruginosaBiofilm structural stability maintenance
Cell communication and differentiation
[31,32]
Capsular polysaccharide, celluloseSalmonella TyphimuriumAdhesion
Environmental survival
[33]
Table 2. Some identified sRNAs with their targets, bacteria, and role in the regulation of target mRNA.
Table 2. Some identified sRNAs with their targets, bacteria, and role in the regulation of target mRNA.
TargetssRNAsBacteriaRegulatory EffectRefs.
AphA, HapRQrr1-4, Qrr1-5V. cholera, V. harveyiActivation[72]
CrcCrcZPseudomonas spp.Repression[73]
CsgDGcvBE. coliRepression[74]
McaSE. coliRepression[75]
OmrA, OmrBE. coliRepression[76]
SdsRS. TyphimuriumActivation[77]
CsgD, YdaMRprAE. coliRepression[78]
CsgD, RpoSArcZS. Typhimurium, E. coliActivation[77,79,80]
CsrACsrB, CsrCYersinia pseudotuberculosis, E. coliRepression[81,82]
PgaAMcaSE. coliActivation[74]
PqsRPhrSP. aeruginosaActivation[83]
RpoSDsrAE. coliActivation[80]
OxySE. coliRepression[80]
RprAE. coliActivation[84]
Table 3. Biofilm-producing pathogenic microbes involved in causing infections.
Table 3. Biofilm-producing pathogenic microbes involved in causing infections.
Pathogenic MicrobesTargeted AreaConsequencesRefs.
Group A streptococciSkinNecrotizing fasciitis, tissue necrosis[108]
Actinobacillus,
Actinomycetemcomitans,
Eikenella corrodens,
Streptococcus mutans,
Prevotella intermedia,
Porphyromonas gingivalis,
Oral spirochetes.
Oral cavityPeriodontal infections, acute inflammation, teeth loosening due to periodontal tissue breakdown, halitosis[109,110]
Streptococcus spp.
Staphylococci (coagulase-negative),
S. aureus,
Enterococcus spp.
Musculoskeletal systemBacterial accumulation on implants and dead bones cause biofilm infections.[111]
Haemophilus influenza,
Streptococcus pneumoniae,
Moraxella catarrhalis.
Middle earOtitis media[112]
S. aureus,
P. aeruginosa.
Lungs (in patients with cystic fibrosis)Mucoviscidosis, lung infections[113]
Table 4. Biofilm-associated foodborne pathogens along with their consequences.
Table 4. Biofilm-associated foodborne pathogens along with their consequences.
OrganismContaminated Food ItemsConsequencesRefs.
Anoxybacillus flavithermusMilk powderReduced acceptability of powdered milk[126]
B. cereusMeat, dairy products, vegetables, and riceVomiting, diarrhea[127,128]
E. coliMeat, vegetables, fruits, and milkHemolytic uremic syndrome, diarrhea[114]
Campylobacter jejuniUnpasteurized milk, animals, poultryVomiting, bloody diarrhea, nausea, fever, and stomach cramps[129]
S. entericaPorcine, bovine, fish, ovine, and poultry meatSepticemia, gastroenteritis[130]
Geobacillus stearothermophilusDairy dried productsEnzymes or acids production resulting in off-flavors[130]
Pseudomonas spp.Meat, vegetables, fruits, and dairy productsBlue discoloration occurrence on fresh cheese[131]
S. aureusDairy products, poultry, eggs, meat, salads, cakes, and pastriesDiarrhea, vomiting[132,133]
L. monocytogenesReady-to-eat products, raw milk, dairy products, Listeriosis in immune-compromised, elderly, and pregnant patients[134]
Pectinatus spp.Brewery environment and beerProduces beer turbid due to the production of sulfur compounds[135]
Table 5. List of some outbreaks caused by biofilm-associated foodborne pathogens.
Table 5. List of some outbreaks caused by biofilm-associated foodborne pathogens.
Region and YearReported CasesResponsible OrganismsFood TypeRef.
South Africa
(2017–2018)
1060L. monocytogenesReady-to-eat meat products[136]
England (2018)34Clostridium perfringensCheese sauce[137]
England (2015)NAE. coli O157:H7Prepacked salad leaves[138]
Massachusetts (2014–2018)1200 per yearSalmonellaNA[139]
England (2016)69CampylobacterRaw milk[140]
Belgium (2013)52S. aureusSeveral foods[141]
China (2010–2014)1040Vibrio parahaemolyticusNA[142]
Europe (2007–2014)6657B. cereusNA[143]
China (2003–2008)9041V. parahaemolyticusMeat and aquatic products[144]
Australia (2001–2010)667L. monocytogenesNA[145]
NA—Not available.
Table 6. Antibiofilm activity of some plant extracts and essential oils.
Table 6. Antibiofilm activity of some plant extracts and essential oils.
Plant Extracts, Compounds or Essential Oils (EOs)Plant SourceBacterial SpeciesInhibition ConcentrationBiofilm Inhibition (%)Ref.
Leaf extractCochlospermum regiumMRSA2000 μg/mL100[169]
5-HydroxymethylfurfuralMusa acuminataP. aeruginosa10 μg/mL83[170]
SyringopicrosideSyringa oblataStreptococcus suis1.28 μg/mL92[171]
XanthohumolHumulus lupulusS. aureus9.8 μg/mL100[172]
SotetsuflavoneCycas media R. BrE. faecalis-60.87–21.74[173]
Lemon grass EOCymbopogon citratusS. aureus250 μg/mL50[174]
P. aeruginosa-
CitralP. aeruginosa0.40 μg/mL
S. aureus>107 μg/mL
EOCinnamomum verumAcinetobacter baumanii, Citrobacter freundii
Corynebacterium striatum, E. coli, Klebsiella spp., S. aureus, Salmonella spp., P. aeruginosa
10 μg/mL97[175]
Thymus vulgaris88
Eugenia caryophyllata91
Ethanol extractAzadirachta indicaMRSA2000 μg/mL43.0[176]
Moringa oleiferaMRSA2000 μg/mL51.4
Murraya koenigiiMRSA2000 μg/mL44.9
Psidium guajavaMRSA2000 μg/mL80
Petroleum ether extractAzadirachta indicaMRSA2000 μg/mL83.8
Moringa oleiferaMRSA2000 μg/mL59.9
Murraya koenigiiMRSA2000 μg/mL63.7
Psidium guajavaMRSA2000 μg/mL62.9
Aqueous extractAcacia niloticaK. pneumoniae13,300 μg/mL59[177]
E.coli13,300 μg/mL63
P. aeruginosa15,000 μg/mL39
Proteus mirabilis16,700 μg/mL49
EOCinnamomum zeylanicumE. coli2000 μg/mL82.76[178]
S. epidermidis83.33
Citrus grandisE. coli2000 μg/mL58.62
S. epidermidis46.67
Citrus hystrixE. coli2000 μg/mL75.86
S. epidermidis83.33
Citrus reticulataE. coli2000 μg/mL82.76
S. epidermidis83.33
Psiadia arguteE. coli2000 μg/mL90
S. epidermidis
Psiadia terebinthinaE. coli2000 μg/mL93.67
S. epidermidis90
Vanilic acid Vaccinium macrocarpon Aiton
(Cranberry)
E. coli23.78 mM100[179]
Protocaterchuic25.95 mM
Catechin55.12, 68.9 mM
Pulp extractEuterpe oleraceaS. aureus250 μg/mL100[180]
Leaf extractJuglans regiaP. aeruginosa16,000 μg/mL60[181]
Leaf extractTetradenia ripariaMRSA-50[182]
Rosmarinus officinalisMRSA30 μg/mL50
ExtractTagetes minutaBacillus sp. Mcn4100 μg/mL50[183]
ExtractTessaria absinthioidesBacillus spp.100 μg/mL66
Staphylococcus sp. Mcr110–50 μg/mL55–62
Sesquiterpene lactonesAcanthospermum hispidumP. aeruginosa0.25–2.5 μg/mL69–77[184]
Table 7. Biofilm inhibition of some plant-based nanoparticles.
Table 7. Biofilm inhibition of some plant-based nanoparticles.
Nanoparticles Plant SpeciesBacteriaInhibition ConcentrationBiofilm Inhibition (%)Ref.
Ag NPsMorinda citrifoliaS. aureus60 μg/mL96[204]
Glochidion lanceolariumP. aeruginosa, E. coli, S. aureus68.9, 12.9, 23.4 μg/mL99[205]
Semecarpus anacardiumP. aeruginosa, E. coli, S. aureus45.5, 23.4, 64.1 μg/mL>99
Bridelia retusaP. aeruginosa, E. coli, S. aureus52.5, 33.8, 32.7 μg/mL
Malus domesticaK. pneumoniae24.6 μg/mL34[206]
Enterobacter aerogenes35.6 μg/mL72
Piper betleP. aeruginosa8 μg/mL78[207]
BER-RHE NPsCoptis chinensis
(Berberine),
Rheum palmatum L. (Rhein)
S. aureus0.1 mmol/mL96[208]
CA-BBR NPsCoptidis rhizome (Berberine)
Cinnamomi cortex (Cinnamic acid)
MRSA0.1 μmol/mL64[209]
Cu NPsCymbopogon citratusE. coli2000 μg/mL49[210]
MRSA33
Crotalaria candicansMRSA1 μg/mL>75[211]
ZnO NPsMyristica fragransE. coli1000 μg/mL24[212]
MRSA1500 μg/mL51
Table 8. Application of some significant nanomaterials for combating biofilms.
Table 8. Application of some significant nanomaterials for combating biofilms.
NanomaterialsTarget OrganismImpact on BiofilmRefs.
CyclodextrinsS. aureus, MRSA, C. albicans,
P. aeruginosa, P. vulgaris,
E. faecalis, E. coli
Adhesion inhibition, biofilm eradication[221,233,234]
DendrimersMRSA, MSSA, E. coli,
K. pneumoniae, P. aeruginosa
Biofilm inhibition[235]
HydrogelsP. aeruginosa, MRSA,
S. aureus, A. baumanii
Biofilm eradication, wound healing[225,236,237]
LiposomesS. aureus, P. gingivalisGrowth inhibition, biofilm formation reduction[238,239]
Polymeric NPsE. coli, S. aureus, S. mutans,
En. Cloacae, P. aeruginosa
Growth inhibition, matrix disruption, and eradication.[240,241]
Stimuli-responsive NPs:
Ag NPs
S. aureus, E. coli, P. aeruginosa, S. flexneri,
K. pneumoniae, S. mutans
C. albicans, S. aureus, E. coli,
P. aeruginosa,
P. aeruginosa, H. pylori,
S. aureus, S. mutans,
M. tuberculosis
Structural alternation, inhibition, oxidative stress.[236,242,243,244]
Au NPsGrowth inhibition, matrix disruption.[245,246,247,248,249]
SPIONsColonization prevention, cell lysis, oxidative stress[250,251,252,253,254]
Solid Lipid NPsS. aureusGrowth inhibition[215,255]
Other inorganic NPsS. aureus, P. aeruginosa,
E. coli, S. epidermidis
Growth inhibition, matrix disruption[205,256,257,258,259]
Table 9. Some of the biofilm control compounds of actinobacteria.
Table 9. Some of the biofilm control compounds of actinobacteria.
Anti-Biofilm CompoundsSourceTarget OrganismBiofilm InhibitionRef.
Actinobacteria
Carotenoid pigmentStreptomyces parvulusC. albicans>50%[273]
Bioactive metabolitesFrankia casuarinae DDNSF-02Candida sp.
Pseudomonas sp.
59–81%
65–80%
[274]
Secondary metabolitesStreptomyces californicus Strain ADR1S. aureus and MRSA90%[275]
Melanin pigmentsNocardiopsis dassonvillei strain JN1, Nocardiopsis sp. JN2Staphylococcus sp.64.20% (JN1)
65.99% (JN2)
[276]
1-hydroxy-1norresistomycin (HNM)Streptomyces variabilisV. cholera
E. coli
S. aureus
92%
96%
93%
[277]
Pyrrolo (1,2-a) pyrazine-1,4-dione hexahydro-3-(2-methylpropyl)Actinomycetes Nocardiopsis sp. GRG 1 (KT235640)E. coli
P. mirabilis
77%
82%
[278]
Actinomycin-DS.parvulusS. aureus
Ruegeria sp.
P. aeruginosa
Micrococcus luteus
53.72%
45.98%
37.12%
22.20%
[279]
Bacterial compounds
Secondary metabolitesStreptomyces (marine sediment)P. mirabilis63–26%[261]
N-acyl homoserine lactone-based QS analogsAqueous extract of Rhizobium sp. NAO1P. aeruginosa77.9%[280]
CarolactonExtract of Sorangium cellulosumStreptococcus oralis,
Streptococcus gordonii,
S. mutans,
A. actinomycetemocomitans
Reduced biofilm formation[281]
CFSClostridium butyricumA. baumannii (MDR strain)Dose-dependent biofilm inhibition[282]
CFSLactobacillus strainsP. aeruginosa0–64%
100% (with L. fermentum L1 and L2)
[283]
CFSLacticasebacillus rhamnosus GGE. coliDose and time-dependent biofilm disruption[284]
Fungal compounds
Diterpenoid sphaeropsidin ADiploidia corticolaP. aeruginosa (MDR strain)
MRSA
62%
53%
[285]
Organic extractsPenicillium sp.P. aeruginosaBiofilm formation reduction by QS inhibition[286]
Vulculic acid, curvulolChaetosphaeronema achilleaeS. aureus DSM 110491.9–96.8%[287]
Thiodiketopiperazine derivativesPhoma sp. GG1F1S. pyogenes
S. aureus
60.7–86%
28–57%
[288]
Crude extractAlternaria alternateP. aeruginosa PAO165.2%[289]
EquisetinFusarium sp. Z10P. aeruginosa PAO158.3%[290]
CFS—Cell-free supernatant.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Asma, S.T.; Imre, K.; Morar, A.; Imre, M.; Acaroz, U.; Shah, S.R.A.; Hussain, S.Z.; Arslan-Acaroz, D.; Istanbullugil, F.R.; Madani, K.; et al. Natural Strategies as Potential Weapons against Bacterial Biofilms. Life 2022, 12, 1618. https://doi.org/10.3390/life12101618

AMA Style

Asma ST, Imre K, Morar A, Imre M, Acaroz U, Shah SRA, Hussain SZ, Arslan-Acaroz D, Istanbullugil FR, Madani K, et al. Natural Strategies as Potential Weapons against Bacterial Biofilms. Life. 2022; 12(10):1618. https://doi.org/10.3390/life12101618

Chicago/Turabian Style

Asma, Syeda Tasmia, Kálmán Imre, Adriana Morar, Mirela Imre, Ulas Acaroz, Syed Rizwan Ali Shah, Syed Zajif Hussain, Damla Arslan-Acaroz, Fatih Ramazan Istanbullugil, Khodir Madani, and et al. 2022. "Natural Strategies as Potential Weapons against Bacterial Biofilms" Life 12, no. 10: 1618. https://doi.org/10.3390/life12101618

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