Next Article in Journal
Design of a Cyclodextrin Bioproduction Process Using Bacillus pseudofirmus and Paenibacillus macerans
Previous Article in Journal
The Landscape of IgA Nephropathy Treatment Strategy: A Pharmacological Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Use of Plant Extracts, Bee-Derived Products, and Probiotic-Related Applications to Fight Multidrug-Resistant Pathogens in the Post-Antibiotic Era

by
António Machado
1,*,
Lizbeth Zamora-Mendoza
2,
Frank Alexis
2,* and
José Miguel Álvarez-Suarez
3,*
1
Laboratorio de Bacteriología, Colegio de Ciencias Biológicas y Ambientales COCIBA, Instituto de Microbiología, Universidad San Francisco de Quito USFQ, Quito 170901, Ecuador
2
Colegio de Ciencias e Ingenierías, Departamento de Ingeniería Química, Universidad San Francisco de Quito USFQ, Quito 170901, Ecuador
3
Colegio de Ciencias e Ingenierías, Departamento de Ingeniería en Alimentos, Universidad San Francisco de Quito USFQ, Quito 170901, Ecuador
*
Authors to whom correspondence should be addressed.
Future Pharmacol. 2023, 3(3), 535-567; https://doi.org/10.3390/futurepharmacol3030034
Submission received: 8 May 2023 / Revised: 5 June 2023 / Accepted: 12 June 2023 / Published: 4 July 2023

Abstract

:
The ‘post-antibiotic’ era is near according to the World Health Organization (WHO). It is well known, due to the work of the scientific community, that drugs (antibiotics, antifungals, and other antimicrobial agents) are continuously becoming less effective, and multidrug-resistant (MDR) pathogens are on the rise. This scenario raises concerns of an impending global infectious disease crisis, wherein a simple opportunistic infection could be deadly for humans. The war against MDR pathogens requires innovation and a multidisciplinary approach. The present study provides comprehensive coverage of relevant topics concerning new antimicrobial drugs; it suggests that a combination of different natural products (such as plant extracts, honey, propolis, prebiotics, probiotics, synbiotics, and postbiotics), together with drug therapy, could be used as an adjuvant in standard treatments, thus allowing drug sensitivity in MDR pathogens to be restored, host immunity to be enhanced, and clinical efficiency to be improved. Currently, new and relevant developments in genomics, transcriptomics, and proteomics are available for research, which could lead to the discovery of new antimicrobial drugs and a new generation of antibiotics and non-antibiotics. However, several areas concerning natural products and their combination with standard drugs remain unclear. In an effort to advance new therapies for humankind, these gaps in the literature need to be addressed.

1. Introduction

The World Health Organization (WHO) expects the ‘post-antibiotic’ era to occur around the year 2050 after evaluating data from 129 member states; every region of the world showed extensive resistance to antimicrobial agents [1,2]. The overuse of antibiotics in several different areas, such as agriculture (to promote livestock growth) and in hospitals (to order standard treatments), has quickly led to the proliferation of drug-resistant bacteria being spread via human travel and poor sanitation practices worldwide [1,3,4,5]. Antimicrobial resistance (AMR) is currently an international issue, and millions of people die every year as a result of opportunistic or primary pathogens that have become resistant due to horizontal gene transfer (HGT) mechanisms and/or biofilm formation [6,7]; this is a multifaceted problem with a catastrophic impact on everyone, including humans, livestock, and the environment [8]. This has led to the estimation that, in 2050, 10 million people will die of infections that cannot be treated because of resistant bacteria and ineffective antibiotics [9].
Currently, healthcare-associated infections (HAI) comprise a main public health concern. These infections usually occur 48 h after hospitalization, although they may also occur after patients are discharged [10]. It is estimated that 7% and 10% of hospitalized patients in developing and developed countries are affected by HAI [10,11], respectively. Moreover, around 3.2 million patients per year are affected by HAI in Europe [10]. The mortality rate and incidence among patients are normally correlated with the patients’ immunological status and geographical region; however, patients in burn units and intensive care units (ICUs), as well as organ transplant receivers and neonates, are the most common hospitalized groups affected by HAI [5,12]. In addition, these infections are also responsible for three out of four lethal cases in neonates in Sub-Saharan Africa and South-East Asia [10]. The most reported HAI are surgical site infections (2–5% incidence rate), catheter-related blood stream infections (12–25% incidence rate), catheter-related urinary tract infections (12% incidence rate), and ventilator-associated pneumonia (9–27% incidence rate) [13]. Currently, the most worrisome global AMRs are the plasmid-mediated spread of carbapenemases (e.g., KPC, NDM, VIM, OXA-48, and OXA-51) and colistin-resistance genes (mcr) in Enterobacteriaceae, Acinetobacter baumannii, and Pseudomonas aeruginosa, as well as the vancomycin resistance gene (vanA) in Enterococcus sp. and Staphylococcus aureus, and the methicillin resistance gene (mecA) in S. aureus [10,13].
This review highlights essential factors contributing to AMR, the epidemiology of the resistant bacteria, and novel alternative therapies that should be developed in subsequent decades to fight the rise of multidrug-resistant (MDR) pathogens. The scientific community and the general public must understand and cooperatively implement the ‘One Health approach’ [8,14]. Neglecting the AMR problem will anticipate the arrival of the ‘post-antibiotic era’; the overuse of antibiotics will increase healthcare costs, morbidity, mortality, and environmental degradation [15,16,17,18]. The lack of new medicines for effective treatments against MDR pathogens, and the emergence of these microorganisms, is a growing global public health concern [19,20]. Despite the critical need for new antimicrobial agents, their rate of development is decreasing [5,12]. Fighting MDR infections calls for a multidisciplinary approach; the present review discusses three alternative antimicrobial drugs and suggests that the combination of different natural products, together with drug therapy, could be used as an adjuvant in standard treatment in order to restore drug sensitivity, enhance host immunity, and improve clinical efficiency. The first section describes the rise of AMR, and the second and third sections discuss natural products with antimicrobial activities, such as plant and honey extracts, respectively. Finally, the fourth section discusses the recent and ongoing developments in microbiome research that are enabling the formulation of new prebiotic, probiotic, and postbiotic products. Therefore, the promising solutions found during the development of new agents are encouraged in the present work. We believe that the success of the long-term battle against MDR pathogens will require new strategies that target other and multiple cellular processes.

2. Rise of MDR Pathogens and Future Trends concerning the Global Infectious Disease Crisis

Understanding the molecular mechanisms underlying antimicrobial (antibiotic or antifungal) resistance is essential, and it requires a deep knowledge of microbial structures and their metabolic functions. Structural and metabolic differences between microorganisms and host cells make it possible to selectively kill the pathogen, or at least inhibit its growth with antimicrobial agents [21], thus allowing the host immune system to eliminate the infection [7]. AMR in pathogens (particularly in bacteria) has emerged as a global challenge since antibiotics were first administered, as it threatens the effectiveness of clinical treatments. In recent decades, there has been an exponential rise in antibiotic resistance-associated factors in microbial communities, most likely driven by the mobility of virulence genes through HGT mechanisms (such as transformation, conjugation, and transduction). Although conjugation, transformation, and transduction are the three primary processes of HGT, six major types of mobile genetic elements (MGEs) have been characterized in the MDR pathogens, such as transposons, gene cassettes, integrons, genomic islands, plasmids, bacteriophages, and integrative conjugative elements (ICEs) [22]. These HGT mechanisms induce genome evolution, and it has caused the rise of different and successful MDR pathogens worldwide [23]. Furthermore, the United States National Institutes of Health (NIH) revealed that around 65% and 80% of all microbial and chronic infections are associated with biofilm formation [24]. The process of biofilm formation consists of many steps, starting with attachment to a living or non-living surface; this leads to the formation of a micro-colony, giving rise to three-dimensional structures, and after maturation, detachment occurs [25,26]. During the formation of biofilm, several species communicate with one another as they employ quorum sensing [27]. In general, biofilms show resistance against the human immune system, as well as against disinfectants and antimicrobials (antibiotics and/or antifungals) [5]. In summary, the understanding of microbial biofilm is important to manage and/or to eradicate biofilm-related diseases. It is believed that biofilms have a great impact on the dissemination of antibiotic resistance as they facilitate HGT mechanisms. Due to the high cell density in the biofilm structure, there is a significant increase in HGT mechanisms. Moreover, the protection given by the extracellular polymeric substances (EPS) of the biofilm by itself intensifies the AMR of the infections with or without the presence of resistant genes in the microbial populations within the biofilm (Figure 1).
Until recently, microorganisms were typically observed as unicellular organisms in the environment; however, currently, it is well known that they prefer to form multicellular communities in nature. As a result, these so-called biofilms are able to endure harsher environments, such as those with a lack of nutrients, natural competitors, and even toxic elements (such as antibiotics) [6,25]. Most microorganisms can live together in biofilms, and the majority of them are known as polymicrobial biofilms, wherein each species and their cells show distinct features when compared with their planktonic form [28]. More specifically, these features include heterogeneity of gene expression, division of roles in the community, presence of persister cells, and enhanced tolerance to antibiotics [29]. In fact, persister cells are dormant and non-dividing cells that exhibit multidrug tolerance and survive treatment from all known antimicrobials [29]. In addition, biofilms are embedded with a highly dense matrix of extracellular polymeric substances (EPS). EPS are complex and potentially diverse polymers produced by the cells within biofilms; they are usually composed of exopolysaccharides, amyloid-like proteins, lipids, and extracellular DNA (eDNA) [26,30]. Despite the intrinsic resistance caused by the presence of EPS surrounding the microbial community within the biofilm, there are also several mechanisms of AMR that have been previously described in other reports; these reports focus on the inactivation or modification of a drug, limiting the drug uptake, drug target modifications, and reducing the active concentration of a drug inside a cell via drug efflux [22,31,32]. The physiological adaptation of microorganisms within a biofilm induces the development of intrinsic resistance, and biofilms are a leading example of resistance to antimicrobial products. Biofilms have been reported to be 100–1000 times more resilient against antimicrobials when compared with equivalent planktonic counterparts [31]. In addition to traversing through the passage of the EPS of the biofilm, the antimicrobial drug needs to enter the microorganism cell membrane at an adequate concentration for a substantial time in order to perform its pharmacological action and produce antimicrobial activity. The efflux pump mechanism is a common mechanism that numerous MDR pathogens use to counter drugs; it extrudes antimicrobial agents faster than it would otherwise [32]. Efflux pumps are proteinaceous, and they are membranal transporters that are able to regulate the microbial cell cytoplasmatic environment; therefore, they can remove toxins, antifungals, and antibiotics. Based on their sequence homologies, the source of energy, substrate binding, and structural components of efflux transporters are usually classified into five prominent families [32]. More specifically, these families are as follows: resistance-nodulation division (RND); adenosine triphosphate (ATP)-binding cassette superfamily (ABC); multidrug and toxic compound extrusion (MATE); major facilitator superfamily (MFS); and small multidrug resistance family (SMR). The two main families of efflux pump proteins in fungi are the ABC and MFS transporters, whereas the RND family is a specific group for Gram-negative bacteria. Finally, the ABC, MATE, MFS, and SMR families are found in Gram-positive and Gram-negative bacteria in significant quantities [32]. However, it is also known that different microorganisms can alter their membrane permeability due to the over- or under-expression of porins, thus controlling the passage of several compounds through the cell membrane; this has been widely reported among Gram-negative pathogens [32]. Finally, another well-known AMR mechanism is the degradation of antimicrobial agents or target site modifications using enzymes. In fact, different enzymes can remove or add a specific moiety to the antibiotic molecule or target site, thus causing a successful mutation on the pathogen which acts against a certain drug [22]. The modification of antimicrobial drugs can be achieved via numerous biochemical reactions that are catalyzed by enzymes; they involve phosphorylation (e.g., chloramphenicol and aminoglycosides), acetylation (e.g., chloramphenicol, aminoglycosides, and streptogramins), and adenylation (e.g., lincosamides and aminoglycosides) [22]. Likewise, target site modifications of several drugs can be succeeded by point mutations in the gene encoding target site, causing an enzymatic change to the target site, and it bypasses the original site [22].
MDR pathogens are one of the most important current threats to public health, and typically, these pathogens are associated with nosocomial and biofilm-related infections [5,33]. In addition, the ineffectiveness of available treatments for such infections has been reported in numerous studies [12,20,34]. Although antibiotics have made it possible to treat deadly infections, the overuse and misuse of antibiotics in recent decades have accelerated the spread of AMR, causing treatments to become ineffective. Currently, at least 700,000 people worldwide die each year due to MDR pathogen-related infections [35]. In 2017, due to the increasing number of AMR reports, the WHO published a list of pathogens, including the pathogens that comprise the acronym ESKAPE (Enterococcus faecium, S. aureus, Klebsiella pneumoniae, A. baumannii, P. aeruginosa, and Enterobacter species), which were given the highest “priority status” as they represent the greatest threat to humans [36]. Several reports note that the rise of these MDR, as well as extensively drug-resistant (XDR) pathogens, render even the most effective drugs ineffective [37,38]. Likewise, in 2022, Cangui and colleagues evaluated the prevalence of biofilms in central venous catheters (CVC) when investigating CVC-related infections in Intensive Care Unit (ICU) patients worldwide. CVC is considered to be one of the deadliest nosocomial or hospital-acquired infections, as S. aureus, coagulase-negative staphylococci, A. baumannii, and P. aeruginosa are the most frequently isolated pathogens [5]. In particular, extended-spectrum β-lactamase (ESBL) and carbapenemase-producing Gram-negative bacteria have emerged as an important therapeutic challenge [4,35,39,40]. Therefore, the development of novel therapies to treat drug-resistant infections, especially those caused by ESKAPE pathogens, is of utmost importance [41,42].

3. Plant Extracts

Historically, plant extracts have been used to treat several bacterial infections in medicine [43]. Several techniques have allowed the extraction and identification of bioactive compounds to recognize the mechanism of action that causes bacteriostatic effects [44,45]. Plant extracts which have antibacterial effects can be obtained from roots, fruits, flowers, stems, leaves, and seeds. Plant extracts are mainly composed of two types of metabolites which are classified as primary and secondary compounds. Primary metabolites are essential compounds for plant survival, whereas secondary metabolites are usually formed in response to plant interactions with the environment [32]. Therefore, primary metabolites are usually products of glycolysis, the shikimate pathway, and the tricarboxylic acid cycle, among other functions which are involved in nutrition and reproduction. Moreover, these metabolites can also act as a precursor to thousands of secondary metabolites that are produced at different steps of the primary metabolic pathways, producing new compounds that facilitate plant adaptations against environmental stress (e.g., bacteria, fungi, insects, disease, injury, temperature, and moisture) [32]. The molecules with the greatest antimicrobial effects are usually secondary metabolites; the most common of these molecules identified in most studies are terpenoids, polyphenols (such as flavonoids, stilbenes, lignans, and phenolic acids), and alkaloids. Many plant terpenoids have found fortuitous uses in medicine [46], and their antimicrobial activity has been attributed to their general membrane disrupting properties [46]. For example, there is evidence that terpenoids of Syzigium cumini exhibit antibacterial activity against methicillin-resistant S. aureus (MRSA) and pathogenic E. coli when minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) assays are used [32]. Likewise, polyphenols’ antimicrobial effects were also documented [32], as were their antioxidant, anti-inflammatory, anticancer, and antihypertensive activities [32,47]. Although the exact mechanism for polyphenols’ antimicrobial action is not fully understood, several polyphenols were reported to exhibit antimicrobial activity against MDR pathogens. Studies postulated that there are different mechanisms at the cellular level, wherein polyphenols can bind to bacterial enzymes via a hydrogen bond, inducing several modifications in terms of cell membrane permeability and cell wall integrity [32,47]. Numerous reports were focused on the most abundant flavonoids, such as flavanols (e.g., quercetin and kaempferol), which demonstrated potent antimicrobial activity against Gram-positive and Gram-negative pathogens, as well as resistant strains [32]. As previously described, the combination of quercetin with amoxicillin exhibited synergistic activity against amoxicillin-resistant Staphylococcus epidermidis isolates [47]. Bryophyllum pinnatum extract revealed that kaempferol, and its derivatives, exhibited significant antimicrobial activity against several bacterial and fungal pathogens, including antibiotic-resistant S. aureus and P. aeruginosa, as well as Candida species and Cryptococcus neoformans [48]. The kaempferol-mediated inhibition of the NorA efflux pump was postulated to be an action mechanism against S. aureus [49]. Finally, alkaloids are organic nitrogenous compounds that are structurally diverse, and their antimicrobial activity has been reported since the 1940s [32]. The mechanism of alkaloids which works against various microbial pathogens is characterized by efflux pump inhibition [50]. In 2020, Duda-Madej and colleagues demonstrated the antibacterial activity of 18 compounds of the O-alkyl derivatives of naringenin and their oximes; these compounds worked against clinical isolates of clarithromycin-resistant Helicobacter pylori, vancomycin-resistant Enterococcus faecalis, methicillin-resistant S. aureus, and beta-lactam-resistant A. baumannii and K. pneumoniae. [51]. Of the pathogen group set, the clarithromycin-resistant strain of H. pylori showed the highest susceptibility to most of the 18 compounds. Moreover, when evaluating the synergy between the O-alkyl derivatives/oximes and several antibiotics via the fractional inhibitory concentration index (FICI), the synergy was observed to be potent when used against H. pylori, S. aureus, and E. faecalis [51].
Nonetheless, other types of compounds have been extensively characterized over the last decade. The pure compounds that have been studied most in recent years are andrographolide, borneol, caffeic acid, thymol, citral, quercetin, epigallocatechin gallate, hydroquinone, oridonin, rhodomyrtosone B, and ursolic acid, among others [52]. Depending on the phytochemical compound, the mechanism of action in different bacteria are cell membrane rupture, aerobic metabolism of interference, protein biosynthesis inhibition, DNA segregation, inhibition of respiratory chain complex proteins, damage to cells’ structural integrity, and disruption of metabolic pathways [52]. The interest in plant extracts has resulted in different patents over the last 20 years, as described in Table 1.
The antibacterial agents are usually used as a pharmaceutical product to treat specific infections depending on the bioactive compounds. Applications in medicine are presented in Figure 2. To understand the therapeutic effects, several studies identified the synergy or antagonistic effects between different compounds to isolate the single bioactive compound from the complete extract [62]. The antimicrobial molecules in the propolis extract studied by Grecka et al. [63] are compounded by well-known flavonoids (pinocembrin, chrysin, and galangin) which work in harmony against several microorganisms. Results have shown that they exhibit a synergistic effect against Gram-positive and Gram-negative bacteria. Furthermore, Rybczyńska-Tkaczyk et al. [64] combined natural compounds with antimicrobial effects to create cosmetics with polyphenolic compounds, which may also exhibit antioxidant and anti-inflammatory effects. Moreover, some plant extracts exhibited significant antibacterial effects against biofilms. The efficacy of these extracts revealed their potential as drug candidates for eradicating pathogenic bacteria [65]. Häsler et al. [66] discussed extraction technologies that can be used in biomedical applications. The preservation of the biologically active compounds depends on the extraction parameters. For example, water is a biocompatible solvent that is commonly used for cosmetic and medical purposes. Other environmentally friendly options are deep eutectic liquids, such as methanol and ethanol, which can replace organic solvents. Moreover, the conventional maceration method is easy and gentle, but extraction effectiveness can be limited [67]. For this reason, other advanced approaches are suggested, such as ultrasound- and microwave-assisted extraction, as they have more benefits in terms of obtaining a high extraction yield. Finally, the plant extract formulations may depend on the applications and the properties of the extract.
Several bacterial infections have been treated with antibacterial extracts incorporated into gels, creams, microspheres, nanoparticles, and hydrogels. For instance, Raju and Jose [68] evaluated the efficacy of a novel topical gel of neem extract using microspheres as a drug delivery system; the gel produced excellent antibacterial effects. The topical application of antimicrobials offers greater advantages at the site of infection. Iraqui et al. [69] reported on an antimicrobial gel containing Cassia alata L. The gel exhibited in vivo wound healing potential due to its significant antibacterial and antifungal activity. Similarly, Popova et al. [70] described a synergistic combination of plant extracts and silver nanoparticles in cream; in the in vivo clinical trials, the cream exhibited significant therapeutic effects on skin diseases, such as antimicrobial action and regenerative effects on tissues. Another structure used for extract-controlled delivery is hydrogels, Nowak et al. [71] presented hydrogels loaded with Epilobium angustifolium L. extract; these exhibited antibacterial effects on dermatological diseases. In addition, Piper crocatum was successfully encapsulated in polyvinyl alcohol, it showed antibacterial activity in Gram-positive and Gram-negative bacteria, and it can be applied to biomedical devices. Indeed, oral capsules containing Cola nitida extract were described by Owusu et al. [72], and they can be used in standard dosages for the management of diarrhea. Finally, the methods that use plant extracts improve biocompatibility and extract-controlled delivery, they decrease toxicity, and enhance biological properties.
There are formulation studies available that used animal models, and some studies have started to use extracts to test for clinically relevant effects in humans. For example, Chaerunisaa et al. [73] demonstrated that Cassia fistula extract provided promising antibacterial activities through in vivo tests on female rats, reporting no alterations to the biochemical parameters of the liver and kidneys of the animal models. Moreover, the zebrafish model was used to evaluate the antimicrobial effects of Lemna minor plant extracts; it produced excellent results and the safety of the extracts and treatment of bacterial septicemia in vivo were also evaluated [74]. There are approximately 180 clinical trials (www.clinicaltrials.gov, accessed on 3 May 2023) in the world that have tested plant extracts, and different combinations of those extracts, in order to treat a variety of diseases (including respiratory diseases, joint diseases, gastrointestinal diseases, etc.). However, no clinical trials testing plant extracts in humans to combat bacterial diseases have been conducted. This suggests a lack of safety and efficacy data, which has prevented tests on plant extracts to investigate their ability to fight human bacterial diseases.

4. Honey and Propolis

Since ancient times, bee products have been recognized for the variety of their biological properties, among which, their antimicrobial activity stands out. In fact, the antibacterial activity of honey was an important finding that was first scientifically described in 1892 by the Dutch scientist Van Ketel [75]. However, other bee-related products (such as propolis) have also been recognized for their biological properties and antimicrobial activities [76,77].
In the case of honey, its antimicrobial properties have been specifically associated with two groups of factors known as (i) peroxide-dependent factors and (ii) non-peroxide-dependent factors. The peroxide-dependent factors of honey are precisely related to the content of hydrogen peroxide (H2O2) that accumulates in it. H2O2 is produced in honey due to the action of the glucose oxidase enzyme (produced by the bee) on glucose, which produces gluconic acid and H2O2 as a by-product of this reaction (Figure 3) [78]; this acts as a sporicidal antiseptic that sterilizes honey and endows it with antibacterial properties against various pathogens.
On the other hand, within the non-peroxide components, osmolality stands out [78]. Honey is a supersaturated solution of sugars, which comprise approximately 80% of its composition. Thus, the osmotically active nature of the sugars causes the dehydration of the bacterial cell, and therefore, its death (Figure 4) [78,79]. In addition to its sugar content, other non-peroxide factors are also important, such as the low pH of honey (between 3.2–4.5) which acts as an inhibitor of various pathogenic bacteria. This acidity is caused by the accumulation of the aforementioned gluconic acid, and other organic acids [78]. There are also non-peroxide factors derived from the floral origin of honey, which include methylglyoxal as the main antimicrobial factor in Manuka honey, other minor components of honey such as phenolic compounds (i.e., flavonoids and phenolic acids), and some unknown floral components. In fact, it has been proposed that the floral origin of honey plays a fundamental role, not only in terms of its physicochemical properties, but also in terms of its antimicrobial activity [80]. Phenols, flavonoids, terpenes, and alkaloids are also included in the group of antimicrobial-related compounds [81], wherein flavanols are one of the most abundant flavonoids present in food (such as in honey and propolis). Flavanols are well-known for their potent antimicrobial activity against Gram-positive and Gram-negative pathogens, including resistant strains [32]. Moreover, little is known about the types of alkaloid compounds in the floral origins of numerous honey products. However, in 2021, Jaktaji and Ghalamfarsa evaluated the interactions between three monofloral honeys (Avishan, Gavan, and Konar) and ciprofloxacin against E. coli [82]. This study demonstrated that all three honey–ciprofloxacin combinations reduced the viability of MG1655 and M1 E. coli strains to a greater extent than ciprofloxacin alone. Moreover, the combination of these honeys and the alkaloid extract of Sophora alopecuroides enhanced the anti-pump activity and reduced the oxidative stress response of the E. coli. Recently, Jaktaji and Koochaki evaluated the in vitro activity of honey and the alkaloid extract of Sophora alopecuroides in combination with antibiotics against four biofilm-producing P. aeruginosa isolates [83]. This study revealed the synergistic effect of alkaloid extract honey in combination with ciprofloxacin against all P. aeruginosa isolates, and it showed a significant reduction in terms of antibiotic resistance and expression of the mexA gene [83]. Both studies demonstrated the importance of alkaloids from plant extracts and honeys as sources of antimicrobial agents, and the importance of their combination with standard drugs when working against MDR and biofilm-associated pathogens [82,83].
On the other hand, there are bee-derived factors, such as the bioactive peptide, defensin-1, and other unknown bee components that can pass from the bee to the honey during the honey production process which cause antimicrobial activity [78].
The spectrum of antibacterial activity for honey is broad, both from the point of view of the flora and geographical origin, and the pathogens that have shown susceptibility to this antibacterial activity. Thus, honey of various floral origins, both polyfloral and monofloral, and honey from different geographical origins, have proven to be effective, not only for inhibiting the growth of pathogenic bacteria in the planktonic state, but also for inhibiting bacterial biofilm formation and eradicating preformed biofilm [84]. Among the most studied pathogens are the Gram-positive S. aureus and MRSA, and the Gram-negative P. aeruginosa, K. pneumoniae (including K. pneumoniae carbapenemase, KPC), and E. coli (Table 2).
Propolis is another bee product that has shown important antimicrobial properties [103]. Propolis, also known as bee glue, is a sticky resinous substance that bees collect from living plants during their nectar and pollen-collecting activities. Its composition is complex, formed mainly by vegetable resins (50%), waxes (30%), aromatic and essential oils (10%), pollen (5%), and other organic compounds (5%). This composition is highly influenced by its floral origin, as is its color, which can range from green to reddish to brown [76]. Regarding its antimicrobial activity, it must be considered through two mechanisms of action. (i) The first is related to the direct action on the microorganism. This action is mainly related to the action of the propolis components on the permeability of the cell membrane of microorganisms, the disruption to the membrane potential, and the production of adenosine triphosphate (ATP), as well as the reduction in bacterial mobility [104]. In fact, the antimicrobial activity of propolis has been reported to be more effective on Gram-positive bacteria than Gram-negative bacteria [103]. This has been explained by the typical structure of the outer membrane of Gram-negative bacteria and the production of hydrolytic enzymes that break down the active components of propolis [105]. The second (ii) mechanism is related to its ability to stimulate the immune system; this results in the activation of the body’s natural defenses [104]. Although the antibacterial activity of propolis has been tested in a varied number of microorganisms, there is a group of bacterial strains that has been more widely analyzed in terms of the strains’ susceptibility to propolis extracts. The ten most tested bacteria for their susceptibility to propolis extracts from different geographic origins include E. coli, S. aureus, Salmonella spp., P. aeruginosa, Yersinia enterocolitica, Enterococcus spp., P. mirabilis, K. pneumoniae, S. mutans, and S. epidermidis [103]. However, studies examining the antibacterial activity of propolis do not only focus upon the aforementioned bacterial groups. Analyses concerning the antibacterial activity of propolis cover a wide group of propolis extracts from different regions of the world and a wide range of microorganisms (Table 3).
Polyphenols and terpenoids are the main components of propolis that have been identified as being responsible for propolis’ antimicrobial activity [106]. Their profiles are closely related to resins and balms of floral origin, as well as the climatic conditions and geographical area where the plants used to produce it grow. Thus, in Europe, North America, and Asia (temperate zone), polyphenolic profiles are characterized by high levels of flavonoids (mainly flavones and flavanones) and low levels of phenolic acids, whereas in tropical zones, propolis shows a more complex composition, with prenylated flavonoids, prenylated p-coumaric acids, and lignans, among others [76]. Several of these compounds have been identified as components of propolis and they exhibit a high degree of antimicrobial activity. An example of such a compound is artepillin C (3,5-diprenyl-p-coumarid acid), a prenyl derivative of p-coumaric acid that can be isolated from propolis. Extracts rich in artepillin C showed a high degree of antibacterial activity against MRSA, as well as against anaerobic bacteria such as Porphyromonas gingivalis, where it exhibited an effective bacteriostatic effect [103] Similarly, other prenyl derivatives reported in propolis, such as 3-prenyl-cinnamic acid allyl ester and 2-dimethyl-8-prenylchromene, have also shown similar antimicrobial activities [107]. Moreover, not only are prenyl derivatives responsible for the antimicrobial activity of propolis, but another abundant group in this product has also been found to have similar properties; this is the flavonoid group. Flavonoids represent a group of important polyphenolic components present in propolis, which are closely related to the high functionality of this bee product [76]. This flavonoid group includes chrysin, pinocembrin, apigenin, galangin, kaempferol, kaempferide, quercetin, tectochrysin, pinostrobin, and others [76]. Pinocembrin isolates were shown to be highly effective against Streptococcus sobrinus, S. mutans, S. aureus, E. faecalis, L. monocytogenes, P. aeruginosa, and K. pneumoniae. Furthermore, isolated apigenin was effective against P. aeruginosa, K. pneumoniae, S. Typhimurium, P. mirabilis, and K. aerogenes. Similarly, the synergistic antibacterial effect of apigenin, together with beta-lactam antibiotics, was also observed against MRSA [45]. In addition, apigenin and ceftazidime also exhibited a synergistic antibacterial effect against ceftazidime-resistant E. cloacae [103]. Propolis, as a material composed largely of plant secretions, is a rich source of phenyl acids, such as cinnamic acid and esters. Several studies have reported the antimicrobial activity of cinnamic acid against various microorganisms, such as Aeromonas spp., Vibrio spp., E. coli, L. monocytogenes, Mycobacterium tuberculosis, Bacillus spp., Staphylococcus spp. S. pyogenes, Micrococcus flavus, P. aeruginosa, S. Typhimurium, E. cloacae, and Yersinia ruckeri [103]. However, not only do these constituents separately contribute to the antimicrobial activity of propolis, but their interactions may be another mechanism by which antimicrobial activity may be enhanced. Thus, the ethanolic extract of propolis, which contains high concentrations of kaempferide, artepillin-C, drupanin, and the phenolic acid, p-coumaric acid, showed significant antibacterial activity against S. aureus, Staphylococcus saprophyticus, L. monocytogenes, and E. faecalis [108].
Table 3. Geographical origin and the main pathogens used in the study of the antimicrobial activity of propolis using the MIC assay.
Table 3. Geographical origin and the main pathogens used in the study of the antimicrobial activity of propolis using the MIC assay.
Geographical OriginBacterial Strain
(Gram-Positive)
ReferencesBacterial Strain
(Gram-Negative)
References
AustraliaS. aureus ATCC 25923[109]K. pneumoniae ATCC 13883[109]
BrazilB. subtilis ATCC 6633, Enterococcus spp., E. faecalis ATTC 29212, ATCC 43300 and ESA 553, Micrococcus luteus ATCC 10240, S. aureus ATCC 6538, ATCC 43300, ATCC 25923, SA 10 and ESA 654, S. epidermidis ATCC 12228 and ESA 675, S. mutans, and S. pyogenes[110,111,112,113,114,115]E. coli ATCC 8739, ATCC 25922 and EC06, K. pneumoniae ATCC 4352 and ESA 154, P. mirabilis ATCC 43300 and ESA 37, P. aeruginosa ATCC 25853, ATCC 15442, PA 24 and ESA 22, and Salmonella spp.[111,113,114,115]
BulgariaS. aureus ATCC 209[116]E. coli WF[116]
ChileS. aureus ATCC 25923 and S. pyogenes ISP 364-00[117]E. coli ATCC 25922 and P. aeruginosa ATCC 27853[117]
Czech RepublicS. aureus ATCC 29213, ATCC 25923 and ATCC 977, S. epidermidis ATCC 14990, S. aureus MRSA/NCTC, S. saprophyticus ATCC 15305, S. oralis ATCC 35037, B. subtilis ATCC 6051, Enterococcus spp., S. agalactiae ATTC 27956, S. pneumoniae ATCC 49619, and S. pyogenes ATCC 12344[118]A. baumani, Burkholderia cepacia, E. cloacae ATCC 700323, E. coli O157:H7, H. influenzae ATCC 49747, K. pneumoniae ATCC 700603., P. aeruginosa ATCC 27853, Salmonella spp., Shigella flexneri, and Y. enterocolitica ATCC 9610[118]
GermanyS. aureus ATCC 29213, ATCC 25923 and ATCC 977, S. epidermidis ATCC 14990, S. aureus MRSA/NCTC, S. saprophyticus ATCC 15305, S. oralis ATCC 35037, B. subtilis ATCC 6051, Enterococcus spp., S. agalactiae ATTC 27956, S. pneumoniae ATCC 49619, and S. pyogenes ATCC 12344[118]A. baumani, B. cepacia, E. cloacae ATCC 700323, E. coli O157:H7, H. influenzae ATCC 49747, K. pneumoniae ATCC 700603., P. aeruginosa ATCC 27853, Salmonella spp., S. flexneri, and Y. enterocolitica ATCC 9610[118]
GreeceS. aureus ATCC 25923 and S. epidermidis ATCC 12228[119]E. cloacae ATCC 13047, E. Coli ATCC 25922, P. aeruginosa ATCC 227853, and K. pneumoniae ATCC 13883[119]
IndiaS. aureus ATCC 6538P[120]
IrelandS. aureus ATCC 29213, ATCC 25923 and ATCC 977, S. epidermidis ATCC 14990, S. aureus MRSA/NCTC, S. saprophyticus ATCC 15305, S. oralis ATCC 35037, B. subtilis ATCC 6051, Enterococcus spp., S. agalactiae ATTC 27956, S. oralis, S. pneumoniae ATCC 49619, and S. pyogenes ATCC 12344[118]A. baumani, B. cepacia, E. cloacae ATCC 700323, E. coli O157:H7, H. influenzae ATCC 49747, K. pneumoniae ATCC 700603, P. aerugino-sa ATCC 27853, Salmonella spp., S. flexneri, and Y. enterocolitica ATCC 9610[118]
ItalyCampylobacter jejuni (clinical isolate) and P. aeruginosa P1242[121,122]
KoreaS. mutans ATCC 25175, S. sobrinus ATCC33478, S. mutans KCOM 1088, KCOM 1091, KCOM 1092, KCOM 1095, KCOM 1097, KCOM 1111, KCOM 1112, KCOM 1113, KCOM 1116, KCOM 1117, KCOM 1118, KCOM 1123, KCOM 1124, KCOM 1126, KCOM 1127, KCOM 1128, KCOM 2762,
KCOM 1136, KCOM 1137, KCOM 1139, KCOM 1142, KCOM 1143, KCOM 1145, KCOM 1146, KCOM 1197, KCOM 1200, KCOM 1201, KCOM 1202, KCOM 1203, KCOM 1207, KCOM 1208, KCOM 1209, KCOM 1212, KCOM 1214, KCOM 1217, KCOM 1219, KCOM 1226 (clinical isolates), and S. sobrinus KCOM 1061, KCOM 1150, KCOM 1151, KCOM 1152, KCOM 1153, KCOM 1157, KCOM 1158, KCOM 1159, KCOM 1185, KCOM 1191, KCOM 1193, KCOM 1196, KCOM 1221, KCOM 1228, and KCOM 1218 (clinical isolates)
[123]
MoroccoS. aureus ATCC 6538 and MRSA 2, 15, and 16 (clinical isolates)[124]
OmanS. aureus ATCC 209[116]E. coli WF[116]
PolandS. aureus ATCC 25923 and S. aureus (clinical isolates) [125]
SlovakiaB. cereus WSBC 10530, S. aureus ATCC 25923, S. aureus Z MJ346, S. pyogenes Z M494, E. faecalis Z MJ90, L. monocytogenes Z M58, and L. monocytogenes Z M70[126]E. coli ATCC 11229, E. coli O157:H7 Z MJ128, S. typhimurium ATCC 14028, S. enteritidis Z M138, C. coli ATCC 33559, C. coli 2235, C. coli 3341-05, C. jejuni ATCC 33560, C. jejuni NCTC 11168, C. jejuni 375-06, and C. jejuni 3552[126]
TurkeyS. mutans ATCC 25175, S. aureus 6538-P, S.sobrinus ATCC 33478, S. epidermidis
ATCC 12228, E. faecalis ATCC 29212, and M. luteus ATCC 9341
[127]P. aeruginosa ATCC
27853, E. coli ATCC 11230, S.
Typhimurium CCM 5445, and K. aerogenes ATCC 13048
[127]

5. Bacteria as a Source of Alternative or Complementary Treatments (Prebiotics, Probiotics, Symbiotics, and Postbiotics)

Since ancient times, bacteria and their products have been used to benefit humans, both in terms of health and food [128]. Prebiotics and probiotics constitute the two main research fields concerning bacterial applications for human benefit. Indeed, the term ‘probiotic’ is actually derived from Greek/Latin and means ‘for life’. Prebiotics are characterized by a group of compounds that are metabolized by microbiota which enhance the growth of probiotic bacteria [128,129]. Over the years, a great deal of research has been conducted on probiotics, and therefore, many definitions have been suggested, however, in 2014, the Food and Agriculture Organization (FAO) and the World Health Organization (WHO) of the United Nations defined probiotics as “live microorganisms which when administered in adequate amounts confer a health benefit on the host” [130]. Likewise, the term ‘prebiotic’ has evolved over the years; however, in 2017, it was defined as “a substrate that is selectively utilized by host microorganisms conferring a health benefit” [131]. A prebiotic substrate must neither be hydrolyzed nor absorbed by the host’s mucosal barrier (such as the gastrointestinal tract), and it must be selectively metabolized by one (or a limited number of) potentially beneficial bacteria that reside in the microbiota. Currently, probiotic and prebiotic concepts have expanded due to recent developments and findings in microbiome research. High-throughput sequencing studies [132,133] have allowed us to improve our knowledge of the composition of microbiota and identify additional substances influencing microbial colonization [131].
Generally, the most studied and used probiotic bacteria belong to the genera Lactobacillus and Bifidobacterium, such as Lactobacillus acidophilus, L. rhamnosus, L. johnsonii, L. casei, L. delbrueckii sp. bulgaricus, L. reuteri, L. brevis, L. fermentum, L. plantarum, Bifidobacterium bifidum, B. adolescentis, B. animalis, B. infantis, and B. thermophilum [134,135,136,137,138,139]. However, other bacteria and yeasts have also been recognized for their probiotic properties, such as B. subtilis, Propionibacterium spp., and S. cerevisiae var. Boulardii [140,141,142,143]. The mechanism of action of probiotics relies on their metabolites, the colonization of the host’s barriers (i.e., skin and mucosal epithelia), competition for nutrients, and production of antimicrobial agents [144], such as lactic acid, hydrogen peroxide, bacteriocins, bacteriocin-like proteins, and biosurfactants [145,146,147]. Probiotic bacteria are known to produce small molecular metabolic byproducts that influence the host’s biological functions. Although some studies are trying to explain how probiotics protect the host [134,138,148,149], these modulation and support mechanisms have not been fully explored. Moreover, the majority of prebiotics are usually characterized by short-chain carbohydrates which are polymerized (≥2) and not susceptible to digestion via the host’s intrinsic enzymes (pancreatic, intestinal, and other mucosal enzymes) [150,151,152]. Most compounds classified as prebiotics are fructans, lactulose, xylooligosaccharides (XOS), and mannan oligosaccharides (MOS). Fructans are mainly composed of several units of fructose linked to terminal sucrose, such as inulin, fructooligosaccharides (FOS), and galactooligosaccharides (GOS) [153,154,155]. Usually classified as health supplements or food products, prebiotics promote the growth of certain probiotic or beneficial bacteria (e.g., lactobacilli, propionibacteria, and bifidobacteria) for the host’s well-being, which improve both epithelial and mucosal protection during external stress or contact with primary and opportunistic pathogens. Prebiotics are also able to promote immune system responses, as shown in several studies which have reported beneficial effects in the gut-associated lymphoid tissue (GALT) [154,156,157]. However, over the last decade, findings have also reported on microbiota being found in other places in the host, such as on the skin and pulmonary mucosa, among others [158,159,160,161].
Recent and ongoing developments in microbiome science are creating new frontiers for research on probiotics and prebiotics (Table 4) [162,163]; indeed, the scientific community has shown an increased interest in synbiotics and postbiotics [155,164,165]. Synbiotics are defined by a combination of probiotics and prebiotics; for instance, a suitable prebiotic could enhance the probiotic’s chance of survival and biological activity. The combination of prebiotics and probiotics could also possess synergistic effects that promote the growth of existing beneficial bacteria in the host, as well as improving the survival, implantation, and growth of newly added probiotic species/strains [128,166]. Synbiotics have not been extensively studied [167,168] and fewer human clinical trials have been carried out on the effectiveness of synbiotics [169,170]. In fact, further studies on different combinations working against different MDR pathogens are needed to create accurate formulations and to undertake further evaluations. It is important to mention that individual probiotic species and strains have different prebiotic requirements, and furthermore, different probiotic species and strains are more efficient against certain pathogen-related infections [135,138]. Moreover, the term ‘postbiotics’ is defined as the “preparation of inanimate microorganisms and/or their components that confers a health benefit on the host”; this definition was devised by a scientific panel from the International Scientific Association for Probiotics and Prebiotics (ISAPP) [171]. Recognizing that the term ‘postbiotic’ means “after life” (not “from life”), postbiotics are characterized by the presence of inanimate microbes and/or physiologically active microbial cellular components (such as cell surface fragments, enzymes, and metabolites) that can contribute to the complexity and functionality of the host’s beneficial effects against illness or infectious diseases [171]. These metabolites are recognized as metabolic byproducts that are secreted by the microorganism; they include enzymes, peptides, teichoic or organic acids, polysaccharides, and other compounds. Therefore, postbiotics are derived from probiotic or even non-probiotic cells, thus providing health benefits for the host when applied in adequate quantities and in requisite combinations [172]. However, the health-improving properties or aspects of postbiotics, and their bioactivities, are still unknown or unclear [173]. Currently, it is possible to better understand the advantages of these postbiotic compounds on one’s overall health due to the new -omic analyses (such as genomics, transcriptomics, metabolomics, lipidomics, and proteomics) that are currently being examined in several investigations [174,175].
Human microbiota are considered ”functional organisms” as a complex community of microorganisms coexist on the host’s skin and in mucosal tissues in a healthy balance [134,137]. Microbiota are crucial for the metabolism and immune system regulation, as well as for the prevention of potential pathogen colonization. Thus, an imbalance in human microbiota can be detrimental to the host’s equilibrium and can cause a state of dysbiosis [150]. Hence, studies into the rise of emerging MDR pathogens have demonstrated how the use of prebiotics and probiotics can help fight these virulent and antibiotic-resistant pathogens in the near post-antibiotic era. In 2019, Joshi and colleagues studied Punica granatum peel extract for its quorum-modulatory potential against two different human-pathogenic bacteria, viz. Chromobacterium violaceum and P. aeruginosa; it exhibited notable prebiotic potential by promoting the growth of B. bifidum and L. plantarum probiotic strains [176]. Moreover, the virulent traits of P. aeruginosa, such as hemolytic activity and biofilm formation, were negatively affected by this extract in in vitro assays, and its therapeutic efficiency was confirmed as the nematode, C. elegans, was also more susceptible to lysis by human sera [176]. Alessandri and colleagues analyzed the effect of chitin-glucan (CG), as a biopolymer of A. niger, on one hundred bifidobacterial strains from infant feces and the gastrointestinal tract of adults [177]; the study demonstrated that almost all bifidobacterial species displayed high growth levels in the in vitro assays, in particular, the B. breve 2L isolate. When evaluating the colonization of B. breve 2L in the mammalian gut via CG stimulation, the in vivo Groningen rat model (Rattus norvegicus) exhibited a significant increase in the gut of B. breve 2L, thus enhancing the gut colonization/persistence of this strain and suggesting that CG exerts a species specific modulation of the bifidobacterial population that is harnessed by the rat gut [177]. In 2022, He and colleagues studied ursodeoxycholic acid (UDCA) activity using in vitro assays to examine the growth of the Escherichia coli serotype, O101:H9, which is isolated from dairy calves and the Caco-2 cell line; it exhibited direct antibacterial effects, suppressed proinflammatory effects (such as IL-1β and IL-10 regulation), and reduced damage to the cell’s integrity [178]. In vivo assays used on a specific pathogen-free (SPF) CD-1 neonatal mice model, significant antibacterial effects were also demonstrated, and they helped maintain colonic barrier integrity. In fact, UDCA supplementation attenuated colitis symptoms and recovered colonic short-chain fatty acid (SCFA) production. Through 16S rRNA gene sequencing, microbiotas from UDCA-treated neonatal mice ameliorated colitis symptoms, as evidenced by the successful colonization of bacteria, including Oscillospiraceae, Ruminococcaceae, Lachnospiraceae, and Clostridia_UCG-014, when compared with control and placebo microbiotas [178]. It is important to note that this prebiotic application was successful against an enteroaggregative E. coli (EAEC) and a multidrug-resistant extended-spectrum β-lactamase (ESBL)-producing E. coli isolate. Furthermore, probiotic applications are also evolving due to the isolation of new and more probiotic strains from a diverse set of samples, and more exhaustive in vitro and in vivo studies. For example, Kao and colleagues examined the extracellular electrons transferred from the honey-derived probiotic, B. circulans, which inhibits the human skin pathogen, C. acnes, by injecting the pathogen intradermally into mice ears to induce an inflammatory response [179]. The results showed that the in vitro B. circulans co-culture enhanced electron production and significantly suppressed C. acnes growth. Moreover, in the in vivo assays of the ears of the Crl:CD1(ICR) mice model, the C. acnes and macrophage inflammatory protein 2 (MIP-2) levels suggested that B. circulans-generated electrons affected C. acnes growth and alleviated the resultant inflammatory response [179]. Islam and colleagues also evaluated a new probiotic strain, B. amyloliquefaciens (BA PMC-80), which exhibited significant anti-C. difficile effects in a co-cultured in vitro assay [180]. It also exhibited no toxicity in a subchronic toxic in vivo hamster model; indeed, a reduction in infection severity and delayed death were observed. However, further studies are required to identify the antimicrobial compound produced by BA PMC-80, which would improve the treatment of the C. difficile infection (CDI) hamster model. Lastly, Ishnaiwer and colleagues found two new strains of B. subtilis (CH311 and S3B) and evaluated them against an ESBL-producing E. coli isolate, wherein both probiotic strains reduced ESBL-E. coli titers by 4 log colony-forming units (CFU)/mL in an in vitro model of gut content [181]. However, an in vivo murine model of intestinal colonization showed no reduction in the fecal titers of the ESBL-E. coli strains, CH311 and S3B [181]. Thus, this study emphasizes the importance of in vivo experiments to identify efficient probiotics, and more importantly, to identify probiotic administration procedures that allow the development and improvement of effective delivery systems.
Other promising applications are synbiotics and postbiotics. Cui and colleagues applied P. acidilactici and phthalyl inulin nanoparticles (PINs) to be used against S. Gallinarum via an in vitro cocultivation assay [182]. The antibacterial activity of the symbiotic formulation was the highest among the treated groups (bacteria control and only PINs or probiotics), exhibiting a statistical reduction from log 8 to log 5 CFU/mL. Interestingly, PINs alone did not demonstrate any antibacterial activity, thus highlighting the synergistic inhibitory effect of this synbiotic formulation on this specific foodborne pathogen [182]. Moreover, Hashem and colleagues evaluated the encapsulation efficiency of alginate-CaCl2 nanoparticles to be used against S. cerevisiae, as well as Moringa oleifera leaf extract (MOLE) to be used against multiple foodborne pathogens, including E. coli BA 12296B, S. aureus NCTC 10788, C. albicans ATCC MYA-2876), L. monocytogenes ATCC 19116, and S. Senftenberg ATCC 8400 [183]. The antimicrobial activity test for administered synbiotics uses the agar-well diffusion method, and it revealed significantly greater diameters for the inhibition zones of the nanoencapsulated synbiotic when compared with the nonencapsulated group against all tested pathogenic bacteria and fungi. Furthermore, in vivo assays used on the rabbit model produced no effects in terms of interleukin-l or immunoglobulin G (IgG) and IgA levels. Moreover, nanoencapsulated synbiotics significantly increased the number of beneficial intestinal and cecal microbes (yeast and lactic-acid bacteria) while reducing the number of coliforms and Salmonella sp. Lastly, in vitro gastrointestinal simulation tests revealed the highest protective effect for the survivability of the probiotic, S. cerevisiae, during gastric and intestinal enzymatic digestion [183]. In 2021, Hong and colleagues evaluated a new formulation of phthalyl pullulan nanoparticles (PPNs) to enhance the antimicrobial activity of the probiotic, L. plantarum, in a dysbiosis-induced murine model using the Escherichia coli K99 pathogen. The authors showed that the infection was significantly suppressed using synbiotics, and several well-known beneficial bacteria, such as Lactobacillus and Bifidobacterium, were enriched [167]. Likewise, in 2022, Bandyopadhyay and colleagues assessed the probiotic effect of two new L. lactis subsp. lactis strains against pathogenic and food spoilage bacteria and fungi (such as Bacillus sp., E. faecalis, E. coli, L. monocytogenes, Aspergillus sp., C. albicans, and Fusarium oxysporum among others); these probiotic strains demonstrated good antimicrobial activity against all pathogenic and food spoilage fungi tested in the study [168]. The further in vivo bacterial feeding of these strains for 30 days in Swiss albino mice either individually, or in combination with prebiotic inulin, improved gut colonization and immunoglobulin A (IgA) production levels. Additionally, the continued feeding provided health benefits that were better than the use of a commercial probiotic consortium together with a prebiotic mixture [168].
As previously stated, there are still not many in vivo studies on postbiotics. According to the National Center for Biotechnology Information (NCBI; https://www.ncbi.nlm.nih.gov/ accessed on 27 April 2023), most in vivo studies on postbiotics were published after 2019. Sornsenee and colleagues recently evaluated the in vitro antimicrobial effects, antioxidant activity, and anti-inflammatory effects of 10 lyophilized cell-free supernatants (LCFS) of Lactobacillus isolates from the fermented palm sap of trees from Southern Thailand, which were used against E. coli DMST4212, A. baumannii DMST 2271, S. aureus DMST 2928, and one clinical MRSA isolate [164]. All LCFS exhibited strong antibiofilm activity, they eradicated the biofilms formed by A. baumannii and E. coli, and they reduced the production of nitric oxide in RAW 264.7 cells [164]. Jung and colleagues analyzed the potential beneficial effects (for the host) of the LCFS mixture of L. plantarum KM1, L. plantarum KM2, and B. velezensis KMU01 (1:1:1; vol/vol) in RAW264.7 cells and a C57BL/6N mice model [184]; the study reported a significant reduction in tumor necrosis factor-alpha (TNF-α) levels and an increase in natural killer (NK) cell activation. In addition, the postbiotic mixture was able to modulate the abundance of Bifidobacteria, Lachnospiraceae, and Lactobacillaceae in the gut of the C57BL/6N mice model [184]. Golkar and colleagues also evaluated the immunomodulatory, anti-inflammatory, and antimicrobial activities of three individual postbiotic cold creams using the LCFS of L. fermentum, L. reuteri, and B. subtilis sp. natto to examine wound healing in a Sprague Dawley rat model [185]. Wound healing in animals using all three cold cream formulations exhibited faster recovery times when compared with animals that were given no treatment or only cold cream by itself. After day 4, all three postbiotic cold creams exhibited higher and significantly better wound healing abilities in comparison with the untreated group and the group treated with cold cream without postbiotics (p < 0.0001). The epithelialization process was complete in rats receiving L. reuteri and B. subtilis cold creams, whereas the L. reuteri cold cream inhibited the inflammation process in treated animals. Finally, animals treated with L. reuteri and B. subtilis cold creams did not demonstrate any histological alterations with regard to granulation [185]. Moreover, Puccetti and colleagues demonstrated the applicability of postbiotics via a spray-dried formulation of indole-3-carboxaldehyde (3-IAld) that was used against Aspergillus fumigatus using two cell lines (Beas-2B and Calu-3) and a C57BL/6 mice model in order to assay potential pulmonary toxicity and inflammatory cytokine gene expression, respectively [186]. The results demonstrated dual therapeutic benefits; the formulation can be used as an anti-inflammatory agent to prevent lung inflammation, and it can be used to reduce aspergillosis disease scores when locally delivered into the lungs via inhalable 3-IAld-Man powder [186].
These studies demonstrated the potential of synbiotics and postbiotics for immune system regulation and host-microbiome modulation. However, more in vitro, and especially in vivo, studies are needed to fully understand the impact and outcomes of synbiotics and postbiotics in host-microbiome and immune system responses against MDR pathogens before the post-antibiotic era begins.

6. Conclusions

The imminent ‘post-antibiotic’ era needs alternative therapies, innovation, and a multidisciplinary approach. The present review compiled the most recent studies that use different natural products to achieve this (more specifically, plant extracts, honey, propolis, prebiotics, probiotics, synbiotics, and postbiotics). These products, either alone or together with standard drug therapies, frequently enabled drug sensitivity in MDR pathogens to be restored, and treatments were improved via in vitro and in vivo assays. A new generation of non-antibiotic compounds is needed to fight MDR pathogens. Further studies evaluating these products, using genomics, transcriptomics, and proteomics, are necessary to address the gaps in the literature by identifying specific compounds and their potential to be used against MDR pathogens, thus allowing the development of new therapies.

Author Contributions

Conceptualization, A.M., F.A. and J.M.Á.-S.; methodology, A.M.; validation, A.M., F.A. and J.M.Á.-S.; formal analysis, A.M., F.A., L.Z.-M. and J.M.Á.-S.; investigation, A.M., F.A., L.Z.-M. and J.M.Á.-S.; resources, A.M.; data curation, A.M.; writing—original draft preparation, A.M., F.A., L.Z.-M. and J.M.Á.-S.; writing—review and editing, A.M., F.A. and J.M.Á.-S.; visualization, A.M.; supervision, A.M.; project administration, A.M.; funding acquisition, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Collaboration Grants 2021 to António Machado of the Research Office from Universidad San Francisco de Quito USFQ, under Project ID: 17577 titled “The antibiofilm potential of lactobacilli biosurfactants against multi-drug-resistant pathogens”. The funders had no role in the study design, data collection, analysis, decision to publish, or preparation of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A special recognition is due to all colleagues at the Microbiology Institute of USFQ, COCIBA, El Politécnico, and Research Office of Universidad San Francisco de Quito for their support in this study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Reardon, S. WHO Warns against “post-Antibiotic” Era. Nature 2014, 15, 135–138. [Google Scholar] [CrossRef]
  2. Kwon, J.H.; Powderly, W.G. The Post-Antibiotic Era Is Here. Science 2021, 373, 471. [Google Scholar] [CrossRef] [PubMed]
  3. Vinueza, D.; Ochoa-Herrera, V.; Maurice, L.; Tamayo, E.; Mejía, L.; Tejera, E.; Machado, A. Determining the Microbial and Chemical Contamination in Ecuador’s Main Rivers. Sci. Rep. 2021, 11, 17640. [Google Scholar] [CrossRef] [PubMed]
  4. Montero, L.; Irazabal, J.; Cardenas, P.; Graham, J.P.; Trueba, G. Extended-Spectrum Beta-Lactamase Producing-Escherichia Coli Isolated from Irrigation Waters and Produce in Ecuador. Front. Microbiol. 2021, 12, 709418. [Google Scholar] [CrossRef]
  5. Cangui-Panchi, S.P.; Ñacato-Toapanta, A.L.; Enríquez-Martínez, L.J.; Reyes, J.; Garzon-Chavez, D.; Machado, A. Biofilm-Forming Microorganisms Causing Hospital-Acquired Infections from Intravenous Catheter: A Systematic Review. Curr. Res. Microb. Sci. 2022, 3, 100175. [Google Scholar] [CrossRef]
  6. Abe, K.; Nomura, N.; Suzuki, S. Biofilms: Hot Spots of Horizontal Gene Transfer (HGT) in Aquatic Environments, with a Focus on a New HGT Mechanism. FEMS Microbiol. Ecol. 2020, 96, fiaa031. [Google Scholar] [CrossRef]
  7. Cangui-Panchi, S.P.; Ñacato-Toapanta, A.L.; Enríquez-Martínez, L.J.; Salinas-Delgado, G.A.; Reyes, J.; Garzon-Chavez, D.; Machado, A. Battle Royale: Immune Response on Biofilms—Host-Pathogen Interactions. Curr. Res. Immunol. 2023, 4, 100057. [Google Scholar] [CrossRef]
  8. Chandra, P.; Unnikrishnan, M.K.; Vandana, K.E.; Mukhopadhyay, C.; Dinesh Acharya, U.; Surulivel Rajan, M.; Rajesh, V. Antimicrobial Resistance and the Post Antibiotic Era: Better Late than Never Effort. Expert. Opin. Drug. Saf. 2021, 20, 1375–1390. [Google Scholar] [CrossRef]
  9. Hansson, K.; Brenthel, A. Imagining a Post-Antibiotic Era: A Cultural Analysis of Crisis and Antibiotic Resistance. Med. Hum. 2022, 48, 381–388. [Google Scholar] [CrossRef]
  10. Avershina, E.; Shapovalova, V.; Shipulin, G. Fighting Antibiotic Resistance in Hospital-Acquired Infections: Current State and Emerging Technologies in Disease Prevention, Diagnostics and Therapy. Front. Microbiol. 2021, 12, 2044. [Google Scholar] [CrossRef]
  11. Tacconelli, E.; Carrara, E.; Savoldi, A.; Harbarth, S.; Mendelson, M.; Monnet, D.L.; Pulcini, C.; Kahlmeter, G.; Kluytmans, J.; Carmeli, Y.; et al. Discovery, Research, and Development of New Antibiotics: The WHO Priority List of Antibiotic-Resistant Bacteria and Tuberculosis. Lancet Infect. Dis. 2018, 18, 318–327. [Google Scholar] [CrossRef]
  12. Atiencia-Carrera, M.B.; Cabezas-Mera, F.S.; Tejera, E.; Machado, A. Prevalence of Biofilms in Candida Spp. Bloodstream Infections: A Meta-Analysis. PLoS ONE 2022, 17, e0263522. [Google Scholar] [CrossRef]
  13. Khan, H.A.; Baig, F.K.; Mehboob, R. Nosocomial Infections: Epidemiology, Prevention, Control and Surveillance. Asian Pac. J. Trop. Biomed. 2017, 7, 478–482. [Google Scholar] [CrossRef]
  14. Wang, F.; Fu, Y.H.; Sheng, H.J.; Topp, E.; Jiang, X.; Zhu, Y.G.; Tiedje, J.M. Antibiotic Resistance in the Soil Ecosystem: A One Health Perspective. Curr. Opin. Env. Sci. Health 2021, 20, 100230. [Google Scholar] [CrossRef]
  15. Kraemer, S.A.; Ramachandran, A.; Perron, G.G. Antibiotic Pollution in the Environment: From Microbial Ecology to Public Policy. Microorganisms 2019, 7, 180. [Google Scholar] [CrossRef] [Green Version]
  16. Baker, S.; Thomson, N.; Weill, F.X.; Holt, K.E. Genomic Insights into the Emergence and Spread of Antimicrobial-Resistant Bacterial Pathogens. Science 2018, 360, 733–738. [Google Scholar] [CrossRef] [Green Version]
  17. Miller, E.A.; Ponder, J.B.; Willette, M.; Johnson, T.J.; VanderWaal, K.L. Merging Metagenomics and Spatial Epidemiology To Understand the Distribution of Antimicrobial Resistance Genes from Enterobacteriaceae in Wild Owls. Appl. Environ. Microbiol. 2020, 86, e00571-20. [Google Scholar] [CrossRef]
  18. McEwen, S.A.; Collignon, P.J. Antimicrobial Resistance: A One Health Perspective. Microbiol. Spectr. 2018, 6, 255–260. [Google Scholar] [CrossRef] [Green Version]
  19. Rybak, J.M.; Barker, K.S.; Muñoz, J.F.; Parker, J.E.; Ahmad, S.; Mokaddas, E.; Abdullah, A.; Elhagracy, R.S.; Kelly, S.L.; Cuomo, C.A.; et al. In Vivo Emergence of High-Level Resistance during Treatment Reveals the First Identified Mechanism of Amphotericin B Resistance in Candida Auris. Clin. Microbiol. Infect. 2022, 28, 838–843. [Google Scholar] [CrossRef]
  20. Muñoz-Barreno, A.; Cabezas-Mera, F.; Tejera, E.; Machado, A. Comparative Effectiveness of Treatments for Bacterial Vaginosis: A Network Meta-Analysis. Antibiotics 2021, 10, 978. [Google Scholar] [CrossRef]
  21. Dale-Skinner, J.W.; Bonev, B.B. Molecular Mechanisms of Antibiotic Resistance: The Need for Novel Antimicrobial Therapies. In New Strategies Combating Bacterial Infection; Blackwell Publishing Ltd.: Oxford, UK, 2009; pp. 1–46. ISBN 9783527322060. [Google Scholar]
  22. Chawla, M.; Verma, J.; Gupta, R.; Das, B. Antibiotic Potentiators against Multidrug-Resistant Bacteria: Discovery, Development, and Clinical Relevance. Front. Microbiol. 2022, 13, 887251. [Google Scholar] [CrossRef] [PubMed]
  23. Klemm, E.J.; Wong, V.K.; Dougan, G. Emergence of Dominant Multidrug-Resistant Bacterial Clades: Lessons from History and Whole-Genome Sequencing. Proc. Natl. Acad. Sci. USA 2018, 115, 12872–12877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. 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]
  25. Atiencia-Carrera, M.B.; Cabezas-Mera, F.S.; Vizuete, K.; Debut, A.; Tejera, E.; Machado, A. Evaluation of the Biofilm Life Cycle between Candida Albicans and Candida Tropicalis. Front. Cell. Infect. Microbiol. 2022, 12, 953168. [Google Scholar] [CrossRef]
  26. 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]
  27. Bjarnsholt, T. Introduction to Biofilms. In Biofilm Infections; Bjarnsholt, T., Moser, C., Jensen, P.Ø., Høiby, N., Eds.; Springer: New York, NY, USA, 2011; pp. 1–9. ISBN 9781441960832. [Google Scholar]
  28. Machado, A.; Cerca, N. Influence of Biofilm Formation by Gardnerella Vaginalis and Other Anaerobes on Bacterial Vaginosis. J. Infect. Dis. 2015, 212, 1856–1861. [Google Scholar] [CrossRef] [Green Version]
  29. Lewis, K. Persister Cells, Dormancy and Infectious Disease. Nat. Rev. Microbiol. 2007, 5, 48–56. [Google Scholar] [CrossRef]
  30. Rubini, D.; Banu, S.F.; Nisha, P.; Murugan, R.; Thamotharan, S.; Percino, M.J.; Subramani, P.; Nithyanand, P. Essential Oils from Unexplored Aromatic Plants Quench Biofilm Formation and Virulence of Methicillin Resistant Staphylococcus Aureus. Microb. Pathog. 2018, 122, 162–173. [Google Scholar] [CrossRef]
  31. Simões, M.; Bennett, R.N.; Rosa, E.A.S. Understanding Antimicrobial Activities of Phytochemicals against Multidrug Resistant Bacteria and Biofilms. Nat. Prod. Rep. 2009, 26, 746–757. [Google Scholar] [CrossRef]
  32. Jubair, N.; Rajagopal, M.; Chinnappan, S.; Abdullah, N.B.; Fatima, A. Review on the Antibacterial Mechanism of Plant-Derived Compounds against Multidrug-Resistant Bacteria (MDR). Evid.-Based Complement. Altern. Med. 2021, 2021, 3663315. [Google Scholar] [CrossRef]
  33. van Duin, D.; Paterson, D.L. Multidrug-Resistant Bacteria in the Community: An Update. Infect. Dis. Clin. N. Am. 2020, 34, 709–722. [Google Scholar] [CrossRef]
  34. Serra-Burriel, M.; Keys, M.; Campillo-Artero, C.; Agodi, A.; Barchitta, M.; Gikas, A.; Palos, C.; López-Casasnovas, G. Impact of Multi-Drug Resistant Bacteria on Economic and Clinical Outcomes of Healthcare-Associated Infections in Adults: Systematic Review and Meta-Analysis. PLoS ONE 2020, 15, e0227139. [Google Scholar] [CrossRef]
  35. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2023, 12, 116. [Google Scholar] [CrossRef]
  36. Harbarth, S.; Kahlmeter, G.; Kluytmans, J.; Mendelson, M.; Hospital, G.S.; Town, C.; Africa, S.; Pulcini, C.; Singh, N.; Theuretzbacher, U.; et al. Global Priority List of Antibiotic-Resistant Bacteria to Guide Research, Discovery, and Development of New Antibiotics; World Health Organization: Geneva, Switzerland, 2017. [Google Scholar]
  37. Mulani, M.S.; Kamble, E.E.; Kumkar, S.N.; Tawre, M.S.; Pardesi, K.R. Emerging Strategies to Combat ESKAPE Pathogens in the Era of Antimicrobial Resistance: A Review. Front. Microbiol. 2019, 10, 539. [Google Scholar] [CrossRef]
  38. Pandey, R.; Mishra, S.K.; Shrestha, A. Characterisation of Eskape Pathogens with Special Reference to Multidrug Resistance and Biofilm Production in a Nepalese Hospital. Infect. Drug. Resist. 2021, 14, 2201–2212. [Google Scholar] [CrossRef]
  39. Peto, L.; Fawcett, N.J.; Crook, D.W.; Peto, T.E.A.; Llewelyn, M.J.; Walker, A.S. Selective Culture Enrichment and Sequencing of Feces to Enhance Detection of Antimicrobial Resistance Genes in Third-Generation Cephalosporin Resistant Enterobacteriaceae. PLoS ONE 2019, 14, e0222831. [Google Scholar] [CrossRef]
  40. Cerezales, M.; Ocampo-Sosa, A.A.; Álvarez Montes, L.; Díaz Ríos, C.; Bustamante, Z.; Santos, J.; Martínez-Martínez, L.; Higgins, P.G.; Gallego, L. High Prevalence of Extensively Drug-Resistant Acinetobacter Baumannii at a Children Hospital in Bolivia. Pediatr. Infect. Dis. J. 2018, 37, 1118–1123. [Google Scholar] [CrossRef]
  41. Santos, A.C.C.; Malta, S.M.; Dantas, R.C.C.; Coelho Rocha, N.D.; Ariston de Carvalho Azevedo, V.; Ueira-Vieira, C. Antimicrobial Activity of Supernatants Produced by Bacteria Isolated from Brazilian Stingless Bee’s Larval Food. BMC Microbiol. 2022, 22, 127. [Google Scholar] [CrossRef]
  42. Fernandez-Soto, P.; Celi, D.; Tejera, E.; Alvarez-Suarez, J.M.; Machado, A. Cinnamomum Sp. and Pelargonium Odoratissimum as the Main Contributors to the Antibacterial Activity of the Medicinal Drink Horchata: A Study Based on the Antibacterial and Chemical Analysis of 21 Plants. Molecules 2023, 28, 693. [Google Scholar] [CrossRef]
  43. Qassadi, F.I.; Zhu, Z.; Monaghan, T.M. Plant-Derived Products with Therapeutic Potential against Gastrointestinal Bacteria. Pathogens 2023, 12, 333. [Google Scholar] [CrossRef]
  44. Pancu, D.F.; Scurtu, A.; Macasoi, I.G.; Marti, D.; Mioc, M.; Soica, C.; Coricovac, D.; Horhat, D.; Poenaru, M.; Dehelean, C. Antibiotics: Conventional Therapy and Natural Compounds with Antibacterial Activity-a Pharmaco-Toxicological Screening. Antibiotics 2021, 10, 401. [Google Scholar] [CrossRef] [PubMed]
  45. Zamora-Mendoza, L.; Vispo, S.N.; De Lima, L.; Mora, J.R.; Machado, A.; Alexis, F. Hydrogel for the Controlled Delivery of Bioactive Components from Extracts of Eupatorium Glutinosum Lam. Leaves. Molecules 2023, 28, 1591. [Google Scholar] [CrossRef] [PubMed]
  46. Bergman, M.E.; Davis, B.; Phillips, M.A. Medically Useful Plant Terpenoids: Biosynthesis, Occurrence, and Mechanism of Action. Molecules 2019, 24, 3961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Daglia, M. Polyphenols as Antimicrobial Agents. Curr. Opin. Biotechnol. 2012, 23, 174–181. [Google Scholar] [CrossRef]
  48. Tatsimo, S.J.N.; de Dieu Tamokou, J.; Havyarimana, L.; Csupor, D.; Forgo, P.; Hohmann, J.; Kuiate, J.-R.; Tane, P. Antimicrobial and Antioxidant Activity of Kaempferol Rhamnoside Derivatives from Bryophyllum Pinnatum. BMC Res. Notes 2012, 5, 158. [Google Scholar] [CrossRef] [Green Version]
  49. Holler, J.G.; Christensen, S.B.; Slotved, H.-C.; Rasmussen, H.B.; Guzman, A.; Olsen, C.-E.; Petersen, B.; Molgaard, P. Novel Inhibitory Activity of the Staphylococcus Aureus NorA Efflux Pump by a Kaempferol Rhamnoside Isolated from Persea Lingue Nees. J. Antimicrob. Chemother. 2012, 67, 1138–1144. [Google Scholar] [CrossRef] [Green Version]
  50. Khameneh, B.; Iranshahy, M.; Soheili, V.; Fazly Bazzaz, B.S. Review on Plant Antimicrobials: A Mechanistic Viewpoint. Antimicrob. Resist. Infect. Control. 2019, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  51. Duda-Madej, A.; Kozłowska, J.; Krzyżek, P.; Anioł, M.; Seniuk, A.; Jermakow, K.; Dworniczek, E. Antimicrobial O-Alkyl Derivatives of Naringenin and Their Oximes against Multidrug-Resistant Bacteria. Molecules 2020, 25, 3642. [Google Scholar] [CrossRef]
  52. Chassagne, F.; Samarakoon, T.; Porras, G.; Lyles, J.T.; Dettweiler, M.; Marquez, L.; Salam, A.M.; Shabih, S.; Farrokhi, D.R.; Quave, C.L. A Systematic Review of Plants with Antibacterial Activities: A Taxonomic and Phylogenetic Perspective. Front. Pharm. 2021, 11, 2069. [Google Scholar] [CrossRef]
  53. CN115708794A; Antibacterial Essential Oil. Tongfu Group China Co. Ltd.: Nantong, China, 2023.
  54. CN115844777A; Itching-Relieving, Antibacterial and Anti-Inflammatory Composition Containing Plant Extract and Application Thereof. Anhui Chuntang Pharmaceutical Co. Ltd.: Hefei, China, 2022.
  55. CN115399343B; Plant Composite Antibacterial Agent Containing Peony Extract and Preparation Method and Application Thereof. Shandong Ruiying Pharmaceutical Group Co. Ltd.: Shandong, China, 2022.
  56. KR102424044B1; Antibacterial Hand Sanitizer Composition Containing Plant Extract. Gusta Co. Ltd.: Seoul, Republic of Korea, 2021.
  57. CN112868678A; Plant Antibacterial Mite-Killing Agent, Preparation Method Thereof and Daily Necessities Containing Plant Antibacterial Mite-Killing Agent. Guangdong Demay New Materials Technology Co. Ltd.: Shanghai, China, 2021.
  58. WO2021182661A1; Composition for Improving Antibacterial, Anti-Inflammatory, Antiviral, and Immune Functions, Comprising Extract of Ligularia Stenocephala as Active Ingredient. Dongguk University Gyeongju Campus Industry-Academy Cooperation Foundation: Gyeongju-si, Republic of Korea, 2020.
  59. US20210386074A1; Plant Extract Compositions and Methods of Making and Using the Same. Greenology Products Inc.: Raleigh, NC, USA, 2020.
  60. US9138451B2; Plant Extract Hydrolysates and Antibacterial Product Containing the Same. Bionorica SE: Ghent, NY, USA, 2009.
  61. WO2003035093A1; Antibacterial Composition Comprising Plant Extract. Naturobiotech Co. Ltd.: Suwon, Republic of Korea, 2002.
  62. Vaou, N.; Stavropoulou, E.; Voidarou, C.; Tsakris, Z.; Rozos, G.; Tsigalou, C.; Bezirtzoglou, E. Interactions between Medical Plant-Derived Bioactive Compounds: Focus on Antimicrobial Combination Effects. Antibiotics 2022, 11, 1014. [Google Scholar] [CrossRef]
  63. Grecka, K.; Kuś, P.M.; Okińczyc, P.; Worobo, R.W.; Walkusz, J.; Szweda, P. The Anti-Staphylococcal Potential of Ethanolic Polish Propolis Extracts. Molecules 2019, 24, 1732. [Google Scholar] [CrossRef] [Green Version]
  64. Rybczyńska-Tkaczyk, K.; Grenda, A.; Jakubczyk, A.; Kiersnowska, K.; Bik-Małodzińska, M. Natural Compounds with Antimicrobial Properties in Cosmetics. Pathogens 2023, 12, 320. [Google Scholar] [CrossRef]
  65. Ilieva, Y.; Marinov, T.; Trayanov, I.; Kaleva, M.; Zaharieva, M.M.; Yocheva, L.; Kokanova-Nedialkova, Z.; Najdenski, H.; Nedialkov, P. Outstanding Antibacterial Activity of Hypericum Rochelii—Comparison of the Antimicrobial Effects of Extracts and Fractions from Four Hypericum Species Growing in Bulgaria with a Focus on Prenylated Phloroglucinols. Life 2023, 13, 274. [Google Scholar] [CrossRef]
  66. Häsler Gunnarsdottir, S.; Sommerauer, L.; Schnabel, T.; Oostingh, G.J.; Schuster, A. Antioxidative and Antimicrobial Evaluation of Bark Extracts from Common European Trees in Light of Dermal Applications. Antibiotics 2023, 12, 130. [Google Scholar] [CrossRef]
  67. Aspé, E.; Fernández, K. The Effect of Different Extraction Techniques on Extraction Yield, Total Phenolic, and Anti-Radical Capacity of Extracts from Pinus Radiata Bark. Ind. Crops Prod. 2011, 34, 838–844. [Google Scholar] [CrossRef]
  68. Raju, D.; Jose, J. Development and Evaluation of Novel Topical Gel of Neem Extract for the Treatment of Bacterial Infections. J. Cosmet. Derm. 2019, 18, 1776–1783. [Google Scholar] [CrossRef]
  69. Iraqui, P.; Chakraborty, T.; Das, M.K.; Yadav, R.N.S. Herbal Antimicrobial Gel with Leaf Extract of Cassia Alata L. J. Drug Deliv. Ther. 2019, 9, 82–94. [Google Scholar] [CrossRef]
  70. Popova, T.P.; Ignatov, I.; Petrova, T.E.; Kaleva, M.D.; Huether, F.; Karadzhov, S.D. Antimicrobial Activity In Vitro of Cream from Plant Extracts and Nanosilver, and Clinical Research In Vivo on Veterinary Clinical Cases. Cosmetics 2022, 9, 122. [Google Scholar] [CrossRef]
  71. Nowak, A.; Zagórska-Dziok, M.; Perużyńska, M.; Cybulska, K.; Kucharska, E.; Ossowicz-Rupniewska, P.; Piotrowska, K.; Duchnik, W.; Kucharski, Ł.; Sulikowski, T.; et al. Assessment of the Anti-Inflammatory, Antibacterial and Anti-Aging Properties and Possible Use on the Skin of Hydrogels Containing Epilobium Angustifolium L. Extracts. Front. Pharm. 2022, 13, 896706. [Google Scholar] [CrossRef]
  72. Owusu, F.W.A.; Asare, C.O.; Enstie, P.; Adi-Dako, O.; Yeboah, G.N.; Kumadoh, D.; Tetteh-Annor, A.; Amenuke, E.M.; Karen, M. Formulation and in Vitro Evaluation of Oral Capsules and Suspension from the Ethanolic Extract of Cola Nitida Seeds for the Treatment of Diarrhea. Biomed. Res. Int. 2021, 2021, 6630449. [Google Scholar] [CrossRef]
  73. Chaerunisaa, A.Y.; Susilawati, Y.; Muhaimin, M.; Milanda, T.; Hendriani, R.; Subarnas, A. Antibacterial Activity and Subchronic Toxicity of Cassia Fistula L. Barks in Rats. Toxicol. Rep. 2020, 7, 649–657. [Google Scholar] [CrossRef] [PubMed]
  74. González-Renteria, M.; del Carmen Monroy-Dosta, M.; Guzmán-García, X.; Hernández-Calderas, I.; Ramos-Lopez, Y.M.A. Antibacterial Activity of Lemna Minor Extracts against Pseudomonas Fluorescens and Safety Evaluation in a Zebrafish Model. Saudi J. Biol. Sci. 2020, 27, 3465–3473. [Google Scholar] [CrossRef] [PubMed]
  75. Eteraf-Oskouei, T.; Najafi, M. Traditional and Modern Uses of Natural Honey in Human Diseases: A Review. Iran. J. Basic Med. Sci. 2013, 16, 731. [Google Scholar] [PubMed]
  76. Giampieri, F.; Quiles, J.L.; Cianciosi, D.; Forbes-Hernández, T.Y.; Orantes-Bermejo, F.J.; Alvarez-Suarez, J.M.; Battino, M. Bee Products: An Emblematic Example of Underutilized Sources of Bioactive Compounds. J. Agric. Food Chem. 2022, 70, 6833–6848. [Google Scholar] [CrossRef] [PubMed]
  77. Alvarez-Suarez, J.M. Bee Products—Chemical and Biological Properties; Springer International Publishing: Cham, Switzerland, 2017; ISBN 9783319596891. [Google Scholar]
  78. Scepankova, H.; Saraiva, J.A.; Estevinho, L.M. Honey Health Benefits and Uses in Medicine. In Bee Products—Chemical and Biological Properties; Springer International Publishing: Cham, Switzerland, 2017; pp. 83–96. [Google Scholar]
  79. Proaño, A.; Coello, D.; Villacrés-Granda, I.; Ballesteros, I.; Debut, A.; Vizuete, K.; Brenciani, A.; Alvarez-Suarez, J.M. The Osmotic Action of Sugar Combined with Hydrogen Peroxide and Bee-Derived Antibacterial Peptide Defensin-1 Is Crucial for the Antibiofilm Activity of Eucalyptus Honey. LWT 2021, 136, 110379. [Google Scholar] [CrossRef]
  80. García-Tenesaca, M.; Navarrete, E.S.; Iturralde, G.A.; Villacrés Granda, I.M.; Tejera, E.; Beltrán-Ayala, P.; Giampieri, F.; Battino, M.; Alvarez-Suarez, J.M. Influence of Botanical Origin and Chemical Composition on the Protective Effect against Oxidative Damage and the Capacity to Reduce in Vitro Bacterial Biofilms of Monofloral Honeys from the Andean Region of Ecuador. Int. J. Mol. Sci. 2018, 19, 45. [Google Scholar] [CrossRef] [Green Version]
  81. Brudzynski, K. Honey as an Ecological Reservoir of Antibacterial Compounds Produced by Antagonistic Microbial Interactions in Plant Nectars, Honey and Honey Bee. Antibiotics 2021, 10, 551. [Google Scholar] [CrossRef]
  82. Jaktaji, R.P.; Ghalamfarsa, F. Antibacterial Activity of Honeys and Potential Synergism of Honeys with Antibiotics and Alkaloid Extract of Sophora Alopecuroides Plant against Antibiotic-Resistant Escherichia Coli Mutant. Iran. J. Basic. Med. Sci. 2021, 24, 623–628. [Google Scholar] [CrossRef]
  83. Jaktaji, R.P.; Koochaki, S. In Vitro Activity of Honey, Total Alkaloids of Sophora Alopecuroides and Matrine Alone and in Combination with Antibiotics against Multidrug-Resistant Pseudomonas Aeruginosa Isolates. Lett. Appl. Microbiol. 2022, 75, 70–80. [Google Scholar] [CrossRef]
  84. Stefanis, C.; Stavropoulou, E.; Giorgi, E.; Voidarou, C.; Constantinidis, T.C.; Vrioni, G.; Tsakris, A. Honey’s Antioxidant and Antimicrobial Properties: A Bibliometric Study. Antioxidants 2023, 12, 414. [Google Scholar] [CrossRef]
  85. Roshan, N.; Rippers, T.; Locher, C.; Hammer, K.A. Antibacterial Activity and Chemical Characteristics of Several Western Australian Honeys Compared to Manuka Honey and Pasture Honey. Arch. Microbiol. 2017, 199, 347–355. [Google Scholar] [CrossRef]
  86. Lusby, P.E.; Coombes, A.L.; Wilkinson, J.M. Bactericidal Activity of Different Honeys against Pathogenic Bacteria. Arch. Med. Res. 2005, 36, 464–467. [Google Scholar] [CrossRef]
  87. Zapata-Vahos, I.C.; Henao-Rojas, J.C.; Yepes-Betancur, D.P.; Marín-Henao, D.; Giraldo Sánchez, C.E.; Calvo-Cardona, S.J.; David, D.; Quijano-Abril, M. Physicochemical Parameters, Antioxidant Capacity, and Antimicrobial Activity of Honeys from Tropical Forests of Colombia: Apis Mellifera and Melipona Eburnea. Foods 2023, 12, 1001. [Google Scholar] [CrossRef]
  88. Alvarez-Suarez, J.M.; Tulipani, S.; Díaz, D.; Estevez, Y.; Romandini, S.; Giampieri, F.; Damiani, E.; Astolfi, P.; Bompadre, S.; Battino, M. Antioxidant and Antimicrobial Capacity of Several Monofloral Cuban Honeys and Their Correlation with Color, Polyphenol Content and Other Chemical Compounds. Food Chem. Toxicol. 2010, 48, 2490–2499. [Google Scholar] [CrossRef]
  89. Alvarez-Suarez, J.M.; Giampieri, F.; Brenciani, A.; Mazzoni, L.; Gasparrini, M.; González-Paramás, A.M.; Santos-Buelga, C.; Morroni, G.; Simoni, S.; Forbes-Hernández, T.Y.; et al. Apis Mellifera vs Melipona Beecheii Cuban Polifloral Honeys: A Comparison Based on Their Physicochemical Parameters, Chemical Composition and Biological Properties. LWT 2018, 87, 272–279. [Google Scholar] [CrossRef]
  90. Morroni, G.; Alvarez-Suarez, J.M.; Brenciani, A.; Simoni, S.; Fioriti, S.; Pugnaloni, A.; Giampieri, F.; Mazzoni, L.; Gasparrini, M.; Marini, E.; et al. Comparison of the Antimicrobial Activities of Four Honeys from Three Countries (New Zealand, Cuba, and Kenya). Front. Microbiol. 2018, 9, 1378. [Google Scholar] [CrossRef] [Green Version]
  91. Valdés-Silverio, L.A.; Iturralde, G.; García-Tenesaca, M.; Paredes-Moreta, J.; Narváez-Narváez, D.A.; Rojas-Carrillo, M.; Tejera, E.; Beltrán-Ayala, P.; Giampieri, F.; Alvarez-Suarez, J.M. Physicochemical Parameters, Chemical Composition, Antioxidant Capacity, Microbial Contamination and Antimicrobial Activity of Eucalyptus Honey from the Andean Region of Ecuador. J. Apic. Res. 2018, 57, 382–394. [Google Scholar] [CrossRef]
  92. Villacrés-Granda, I.; Proaño, A.; Coello, D.; Debut, A.; Vizuete, K.; Ballesteros, I.; Granda-Albuja, G.; Rosero-Mayanquer, H.; Battino, M.; Giampieri, F.; et al. Effect of Thermal Liquefaction on Quality, Chemical Composition and Antibiofilm Activity against Multiresistant Human Pathogens of Crystallized Eucalyptus Honey. Food Chem. 2021, 365, 130519. [Google Scholar] [CrossRef]
  93. Leyva-Jimenez, F.J.; Lozano-Sanchez, J.; Borras-Linares, I.; de la Luz Cadiz-Gurrea, M.; Mahmoodi-Khaledi, E. Potential Antimicrobial Activity of Honey Phenolic Compounds against Gram Positive and Gram Negative Bacteria. LWT 2019, 101, 236–245. [Google Scholar] [CrossRef]
  94. Coniglio, M.A.; Laganà, P.; Faro, G.; Marranzano, M. Antimicrobial Potential of Sicilian Honeys against Staphylococcus Aureus and Pseudomonas Aeruginosa. J. AOAC Int. 2018, 101, 956–959. [Google Scholar] [CrossRef]
  95. Aumeeruddy, M.Z.; Aumeeruddy-Elalfi, Z.; Neetoo, H.; Zengin, G.; Blom van Staden, A.; Fibrich, B.; Lambrechts, I.A.; Rademan, S.; Szuman, K.M.; Lall, N.; et al. Pharmacological Activities, Chemical Profile, and Physicochemical Properties of Raw and Commercial Honey. Biocatal. Agric. Biotechnol. 2019, 18, 101005. [Google Scholar] [CrossRef]
  96. Lu, J.; Turnbull, L.; Burke, C.M.; Liu, M.; Carter, D.A.; Schlothauer, R.C.; Whitchurch, C.B.; Harry, E.J. Manuka-Type Honeys Can Eradicate Biofilms Produced by Staphylococcus Aureus Strains with Different Biofilm-Forming Abilities. PeerJ 2014, 2, e326. [Google Scholar] [CrossRef] [PubMed]
  97. Ejaz, H.; Sultan, M.; Qamar, M.U.; Junaid, K.; Rasool, N.; Alanazi, A.; Alruways, M.W.; Mazhari, B.B.Z.; Alruwaili, Y.; Bukhari, S.N.A.; et al. Antibacterial Efficacy of Indigenous Pakistani Honey against Extensively Drug-Resistant Clinical Isolates of Salmonella Enterica Serovar Typhi: An Alternative Option to Combat Antimicrobial Resistance. BMC Complement. Med. 2023, 23, 42. [Google Scholar] [CrossRef] [PubMed]
  98. Kunat-Budzyńska, M.; Rysiak, A.; Wiater, A.; Grąz, M.; Andrejko, M.; Budzyński, M.; Bryś, M.S.; Sudziński, M.; Tomczyk, M.; Gancarz, M.; et al. Chemical Composition and Antimicrobial Activity of New Honey Varietals. Int. J. Env. Res. Public Health 2023, 20, 2458. [Google Scholar] [CrossRef] [PubMed]
  99. Majtan, J.; Bohova, J.; Horniackova, M.; Klaudiny, J.; Majtan, V. Anti-Biofilm Effects of Honey against Wound Pathogens Proteus mirabilis and Enterobacter cloacae. Phytother. Res. 2014, 28, 69–75. [Google Scholar] [CrossRef] [PubMed]
  100. Bucekova, M.; Jardekova, L.; Juricova, V.; Bugarova, V.; Di Marco, G.; Gismondi, A.; Leonardi, D.; Farkasovska, J.; Godocikova, J.; Laho, M.; et al. Antibacterial Activity of Different Blossom Honeys: New Findings. Molecules 2019, 24, 1573. [Google Scholar] [CrossRef] [Green Version]
  101. 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. 2019, 67, 688–698. [Google Scholar] [CrossRef] [Green Version]
  102. Kolayli, S.; Kazaz, G.; Özkök, A.; Keskin, M.; Kara, Y.; Demir Kanbur, E.; Ertürk, Ö. The Phenolic Composition, Aroma Compounds, Physicochemical and Antimicrobial Properties of Nigella sativa L. (Black Cumin) Honey. Eur. Food Res. Technol. 2023, 249, 653–664. [Google Scholar] [CrossRef]
  103. Przybyłek, I.; Karpiński, T.M. Antibacterial Properties of Propolis. Molecules 2019, 24, 2047. [Google Scholar] [CrossRef] [Green Version]
  104. Sforcin, J.M.; Bankova, V. Propolis: Is There a Potential for the Development of New Drugs? J. Ethnopharmacol. 2011, 133, 253–260. [Google Scholar] [CrossRef]
  105. Sforcin, J.M. Biological Properties and Therapeutic Applications of Propolis. Phytother. Res. 2016, 30, 894–905. [Google Scholar] [CrossRef]
  106. Pimenta, H.C.; Violante, I.M.P.; de Musis, C.R.; Borges, Á.H.; Aranha, A.M.F. In Vitro Effectiveness of Brazilian Brown Propolis against Enterococcus Faecalis. Braz. Oral. Res. 2015, 29, 1–6. [Google Scholar] [CrossRef] [Green Version]
  107. Viuda-Martos, M.; Ruiz-Navajas, Y.; Fernández-López, J.; Pérez-Álvarez, J.A. Functional Properties of Honey, Propolis, and Royal Jelly. J. Food Sci. 2008, 73, R117–R124. [Google Scholar] [CrossRef]
  108. Seibert, J.B.; Bautista-Silva, J.P.; Amparo, T.R.; Petit, A.; Pervier, P.; dos Santos Almeida, J.C.; Azevedo, M.C.; Silveira, B.M.; Brandão, G.C.; de Souza, G.H.B.; et al. Development of Propolis Nanoemulsion with Antioxidant and Antimicrobial Activity for Use as a Potential Natural Preservative. Food Chem. 2019, 287, 61–67. [Google Scholar] [CrossRef]
  109. Massaro, C.F.; Simpson, J.B.; Powell, D.; Brooks, P. Chemical Composition and Antimicrobial Activity of Honeybee (Apis Mellifera Ligustica) Propolis from Subtropical Eastern Australia. Sci. Nat. 2015, 102, 68. [Google Scholar] [CrossRef]
  110. Bittencourt, M.L.F.; Ribeiro, P.R.; Franco, R.L.P.; Hilhorst, H.W.M.; de Castro, R.D.; Fernandez, L.G. Metabolite Profiling, Antioxidant and Antibacterial Activities of Brazilian Propolis: Use of Correlation and Multivariate Analyses to Identify Potential Bioactive Compounds. Food Res. Int. 2015, 76, 449–457. [Google Scholar] [CrossRef] [Green Version]
  111. Moncla, B.J.; Guevara, P.W.; Wallace, J.A.; Marcucci, M.C.; Nor, J.E.; Bretz, W.A. The Inhibitory Activity of Typified Propolis against. Enterococcus Species. Z. Nat. C 2012, 67, 249–256. [Google Scholar] [CrossRef]
  112. Schmidt, E.M.; Stock, D.; Chada, F.J.G.; Finger, D.; Christine Helena Frankland Sawaya, A.; Eberlin, M.N.; Felsner, M.L.; Quináia, S.P.; Monteiro, M.C.; Torres, Y.R. A Comparison between Characterization and Biological Properties of Brazilian Fresh and Aged Propolis. Biomed. Res. Int. 2014, 2014, 257617. [Google Scholar] [CrossRef] [Green Version]
  113. Campos, J.F.; dos Santos, U.P.; dos Santos da Rocha, P.; Damião, M.J.; Balestieri, J.B.P.; Cardoso, C.A.L.; Paredes-Gamero, E.J.; Estevinho, L.M.; de Picoli Souza, K.; dos Santos, E.L. Antimicrobial, Antioxidant, Anti-Inflammatory, and Cytotoxic Activities of Propolis from the Stingless Bee Tetragonisca fiebrigi (Jataí). Evid.-Based Complement. Altern. Med. 2015, 2015, 296186. [Google Scholar] [CrossRef] [Green Version]
  114. Regueira, M.S.; Tintino, S.R.; da Silva, A.R.P.; do Socorro Costa, M.; Boligon, A.A.; Matias, E.F.F.; de Queiroz Balbino, V.; Menezes, I.R.A.; Melo Coutinho, H.D. Seasonal Variation of Brazilian Red Propolis: Antibacterial Activity, Synergistic Effect and Phytochemical Screening. Food Chem. Toxicol. 2017, 107, 572–580. [Google Scholar] [CrossRef]
  115. Dantas Silva, R.P.; Machado, B.A.S.; de Abreu Barreto, G.; Costa, S.S.; Andrade, L.N.; Amaral, R.G.; Carvalho, A.A.; Padilha, F.F.; Barbosa, J.D.V.; Umsza-Guez, M.A. Antioxidant, Antimicrobial, Antiparasitic, and Cytotoxic Properties of Various Brazilian Propolis Extracts. PLoS ONE 2017, 12, e0172585. [Google Scholar] [CrossRef] [PubMed]
  116. Popova, M.; Dimitrova, R.; Al-Lawati, H.T.; Tsvetkova, I.; Najdenski, H.; Bankova, V. Omani Propolis: Chemical Profiling, Antibacterial Activity and New Propolis Plant Sources. Chem. Cent. J. 2013, 7, 158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Bridi, R.; Montenegro, G.; Nuñez-Quijada, G.; Giordano, A.; Fernanda Morán-Romero, M.; Jara-Pezoa, I.; Speisky, H.; Atala, E.; López-Alarcón, C. International Regulations of Propolis Quality: Required Assays Do Not Necessarily Reflect Their Polyphenolic-Related In Vitro Activities. J. Food Sci. 2015, 80, C1188–C1195. [Google Scholar] [CrossRef] [PubMed]
  118. AL-Ani, I.; Zimmermann, S.; Reichling, J.; Wink, M. Antimicrobial Activities of European Propolis Collected from Various Geographic Origins Alone and in Combination with Antibiotics. Medicines 2018, 5, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Popova, M.P.; Chinou, I.B.; Marekov, I.N.; Bankova, V.S. Terpenes with Antimicrobial Activity from Cretan Propolis. Phytochemistry 2009, 70, 1262–1271. [Google Scholar] [CrossRef]
  120. Patel, J.; Ketkar, S.; Patil, S.; Fearnley, J.; Mahadik, K.R.; Paradkar, A.R. Potentiating Antimicrobial Efficacy of Propolis through Niosomal-Based System for Administration. Integr. Med. Res. 2015, 4, 94–101. [Google Scholar] [CrossRef] [Green Version]
  121. De Marco, S.; Piccioni, M.; Pagiotti, R.; Pietrella, D. Antibiofilm and Antioxidant Activity of Propolis and Bud Poplar Resins versus Pseudomonas Aeruginosa. Evid.-Based Complement. Altern. Med. 2017, 2017, 5163575. [Google Scholar] [CrossRef] [Green Version]
  122. Campana, R.; Patrone, V.; Franzini, I.T.M.; Diamantini, G.; Vittoria, E.; Baffone, W. Antimicrobial Activity of Two Propolis Samples against Human Campylobacter jejuni. J. Med. Food 2009, 12, 1050–1056. [Google Scholar] [CrossRef]
  123. Kim, M.J.; Kim, C.S.; Kim, B.-H.; Ro, S.-B.; Lim, Y.K.; Park, S.-N.; Cho, E.; Ko, J.-H.; Kwon, S.-S.; Ko, Y.-M.; et al. Antimicrobial Effect of Korean Propolis against the Mutans Streptococci Isolated from Korean. J. Microbiol. 2011, 49, 161–164. [Google Scholar] [CrossRef]
  124. El-Guendouz, S.; Aazza, S.; Lyoussi, B.; Bankova, V.; Popova, M.; Neto, L.; Faleiro, M.L.; da Graça Miguel, M. Moroccan Propolis: A Natural Antioxidant, Antibacterial, and Antibiofilm against Staphylococcus aureus with No Induction of Resistance after Continuous Exposure. Evid.-Based Complement. Altern. Med. 2018, 2018, 9759240. [Google Scholar] [CrossRef] [Green Version]
  125. Wojtyczka, R.; Dziedzic, A.; Idzik, D.; Kępa, M.; Kubina, R.; Kabała-Dzik, A.; Smoleń-Dzirba, J.; Stojko, J.; Sajewicz, M.; Wąsik, T. Susceptibility of Staphylococcus Aureus Clinical Isolates to Propolis Extract Alone or in Combination with Antimicrobial Drugs. Molecules 2013, 18, 9623–9640. [Google Scholar] [CrossRef] [Green Version]
  126. Mavri, A.; Abramovič, H.; Polak, T.; Bertoncelj, J.; Jamnik, P.; Smole Možina, S.; Jeršek, B. Chemical Properties and Antioxidant and Antimicrobial Activities of Slovenian Propolis. Chem. Biodivers. 2012, 9, 1545–1558. [Google Scholar] [CrossRef]
  127. Uzel, A.; Sorkun, K.; Önçağ, Ö.; Çoğulu, D.; Gençay, Ö.; Salïh, B. Chemical Compositions and Antimicrobial Activities of Four Different Anatolian Propolis Samples. Microbiol. Res. 2005, 160, 189–195. [Google Scholar] [CrossRef]
  128. Ötleş, S. Probiotics and Prebiotics in Food, Nutrition and Health; Taylor & Francis Group: Boca Raton, FL, USA, 2014; ISBN 9781466586246. [Google Scholar]
  129. de Melo, F.H.C.; Menezes, F.N.D.D.; de Sousa, J.M.B.; dos Santos Lima, M.; da Silva Campelo Borges, G.; de Souza, E.L.; Magnani, M. Prebiotic Activity of Monofloral Honeys Produced by Stingless Bees in the Semi-Arid Region of Brazilian Northeastern toward Lactobacillus Acidophilus LA-05 and Bifidobacterium Lactis BB-12. Food Res. Int. 2020, 128, 108809. [Google Scholar] [CrossRef]
  130. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.; Salminen, S.; et al. Expert Consensus Document: The International Scientific Association for Probiotics and Prebiotics Consensus Statement on the Scope and Appropriate Use of the Term Probiotic. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 506–514. [Google Scholar] [CrossRef] [Green Version]
  131. Gibson, G.R.; Hutkins, R.; Sanders, M.E.; Prescott, S.L.; Reimer, R.A.; Salminen, S.J.; Scott, K.; Stanton, C.; Swanson, K.S.; Cani, P.D.; et al. Expert Consensus Document: The International Scientific Association for Probiotics and Prebiotics (ISAPP) Consensus Statement on the Definition and Scope of Prebiotics. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 491–502. [Google Scholar] [CrossRef] [Green Version]
  132. Martin, D.H. The Microbiota of the Vagina and Its Influence on Women’s Health and Disease. Am. J. Med. Sci. 2012, 343, 2–9. [Google Scholar] [CrossRef] [Green Version]
  133. Suez, J.; Zmora, N.; Elinav, E. Probiotics in the Next-Generation Sequencing Era. Gut Microbes 2020, 11, 77–93. [Google Scholar] [CrossRef]
  134. Wieërs, G.; Belkhir, L.; Enaud, R.; Leclercq, S.; de Foy, J.-M.P.; Dequenne, I.; de Timary, P.; Cani, P.D. How Probiotics Affect the Microbiota. Front. Cell. Infect. Microbiol. 2020, 9, 454. [Google Scholar] [CrossRef] [Green Version]
  135. Abdou, A.M.; Hedia, R.H.; Omara, S.T.; Mahmoud, M.A.E.F.; Kandil, M.M.; Bakry, M.A. Interspecies Comparison of Probiotics Isolated from Different Animals. Vet. World 2018, 11, 227–230. [Google Scholar] [CrossRef] [Green Version]
  136. Zangl, I.; Pap, I.J.; Aspöck, C.; Schüller, C. The Role of Lactobacillus Species in the Control of Candida via Biotrophic Interactions. Microb. Cell 2020, 7, 1–14. [Google Scholar] [CrossRef] [PubMed]
  137. Pacha-Herrera, D.; Erazo-Garcia, M.P.; Cueva, D.F.; Orellana, M.; Borja-Serrano, P.; Arboleda, C.; Tejera, E.; Machado, A. Clustering Analysis of the Multi-Microbial Consortium by Lactobacillus Species against Vaginal Dysbiosis Among Ecuadorian Women. Front. Cell. Infect. Microbiol. 2022, 12, 863208. [Google Scholar] [CrossRef] [PubMed]
  138. Rodríguez-Arias, R.J.; Guachi-Álvarez, B.O.; Montalvo-Vivero, D.E.; Machado, A. Lactobacilli Displacement and Candida Albicans Inhibition on Initial Adhesion Assays: A Probiotic Analysis. BMC Res. Notes 2022, 15, 239. [Google Scholar] [CrossRef] [PubMed]
  139. Chew, S.Y.; Cheah, Y.K.; Seow, H.F.; Sandai, D.; Than, L.T.L. Probiotic Lactobacillus Rhamnosus GR-1 and Lactobacillus Reuteri RC-14 Exhibit Strong Antifungal Effects against Vulvovaginal Candidiasis-Causing Candida Glabrata Isolates. J. Appl. Microbiol. 2015, 118, 1180–1190. [Google Scholar] [CrossRef] [Green Version]
  140. Chelliah, R.; Kim, E.J.; Daliri, E.B.M.; Antony, U.; Oh, D.H. In Vitro Probiotic Evaluation of Saccharomyces boulardii with Antimicrobial Spectrum in a Caenorhabditis elegans Model. Foods 2021, 10, 1428. [Google Scholar] [CrossRef]
  141. Zulkhairi Amin, F.A.; Sabri, S.; Ismail, M.; Chan, K.W.; Ismail, N.; Mohd Esa, N.; Mohd Lila, M.A.; Zawawi, N. Probiotic Properties of Bacillus Strains Isolated from Stingless Bee (Heterotrigona itama) Honey Collected across Malaysia. Int. J. Environ. Res. Public Health 2019, 17, 278. [Google Scholar] [CrossRef] [Green Version]
  142. Dyshlyuk, L.S.; Milentyeva, I.S.; Asyakina, L.K.; Ostroumov, L.A.; Osintsev, A.M.; Pozdnyakova, A.V. Using Bifidobacterium and Propionibacterium Strains in Probiotic Consortia to Normalize the Gastrointestinal Tract. Braz. J. Biol. 2022, 84, 1–17. [Google Scholar] [CrossRef]
  143. Luiz, F.; Do Carmo, R.; Rabah, H.; Fernandes Cordeiro, B.; Da Silva, H.S.; Pessoa, R.M.; Odília, S.; Fernandes, A.; Cardoso, V.N.; Gagnaire, V.; et al. Probiotic Propionibacterium Freudenreichii Requires SlpB Protein to Mitigate Mucositis Induced by Chemotherapy. Oncotarget 2020, 10, 7198–7219. [Google Scholar]
  144. Di Cerbo, A.; Palmieri, B.; Aponte, M.; Morales-Medina, J.C.; Iannitti, T. Mechanisms and Therapeutic Effectiveness of Lactobacilli. J. Clin. Pathol. 2016, 69, 187–203. [Google Scholar] [CrossRef] [Green Version]
  145. Hefzy, E.M.; Khalil, M.A.F.; Ibrahim Amin, A.A.; Ashour, H.M.; Abdelaliem, Y.F. Bacteriocin-like Inhibitory Substances from Probiotics as Therapeutic Agents for Candida Vulvovaginitis. Antibiotics 2021, 10, 306. [Google Scholar] [CrossRef]
  146. Dos Santos, C.I.; França, Y.R.; Campos, C.D.L.; Bomfim, M.R.Q.; Melo, B.O.; Holanda, R.A.; Santos, V.L.; Monteiro, S.G.; Moffa, E.B.; Monteiro, A.S.; et al. Antifungal and Antivirulence Activity of Vaginal Lactobacillus Spp. Products against Candida Vaginal Isolates. Pathogens 2019, 8, 150. [Google Scholar] [CrossRef] [Green Version]
  147. Anjana; Tiwari, S.K. Bacteriocin-Producing Probiotic Lactic Acid Bacteria in Controlling Dysbiosis of the Gut Microbiota. Front. Cell. Infect. Microbiol. 2022, 12, 415. [Google Scholar] [CrossRef]
  148. Allonsius, C.N.; van den Broek, M.F.L.; De Boeck, I.; Kiekens, S.; Oerlemans, E.F.M.; Kiekens, F.; Foubert, K.; Vandenheuvel, D.; Cos, P.; Delputte, P.; et al. Interplay between Lactobacillus Rhamnosus GG and Candida and the Involvement of Exopolysaccharides. Microb. Biotechnol. 2017, 10, 1753–1763. [Google Scholar] [CrossRef]
  149. Matsubara, V.H.; Wang, Y.; Bandara, H.M.H.N.; Mayer, M.P.A.; Samaranayake, L.P. Probiotic Lactobacilli Inhibit Early Stages of Candida Albicans Biofilm Development by Reducing Their Growth, Cell Adhesion, and Filamentation. Appl. Microbiol. Biotechnol. 2016, 100, 6415–6426. [Google Scholar] [CrossRef] [Green Version]
  150. Mendling, W. Microbiota of the Human Body. Adv. Exp. Med. Biol. 2016, 902, 83–93. [Google Scholar] [CrossRef]
  151. Carlson, J.L.; Erickson, J.M.; Lloyd, B.B.; Slavin, J.L. Health Effects and Sources of Prebiotic Dietary Fiber. Curr. Dev. Nutr. 2018, 2, nzy005. [Google Scholar] [CrossRef] [Green Version]
  152. Alexander, C.; Swanson, K.S.; Fahey, G.C.; Garleb, K.A. Perspective: Physiologic Importance of Short-Chain Fatty Acids from Nondigestible Carbohydrate Fermentation. Adv. Nutr. 2019, 10, 576–589. [Google Scholar] [CrossRef] [Green Version]
  153. Abouloifa, H.; Khodaei, N.; Rokni, Y.; Karboune, S.; Brasca, M.; D’Hallewin, G.; Salah, R.B.; Saalaoui, E.; Asehraou, A. The Prebiotics (Fructo-Oligosaccharides and Xylo-Oligosaccharides) Modulate the Probiotic Properties of Lactiplantibacillus and Levilactobacillus Strains Isolated from Traditional Fermented Olive. World J. Microbiol. Biotechnol. 2020, 36, 185. [Google Scholar] [CrossRef]
  154. Roupar, D.; Coelho, M.C.; Gonçalves, D.A.; Silva, S.P.; Coelho, E.; Silva, S.; Coimbra, M.A.; Pintado, M.; Teixeira, J.A.; Nobre, C. Evaluation of Microbial-Fructo-Oligosaccharides Metabolism by Human Gut Microbiota Fermentation as Compared to Commercial Inulin-Derived Oligosaccharides. Foods 2022, 11, 954. [Google Scholar] [CrossRef]
  155. Mounir, M.; Ibijbijen, A.; Farih, K.; Rabetafika, H.N.; Razafindralambo, H.L. Synbiotics and Their Antioxidant Properties, Mechanisms, and Benefits on Human and Animal Health: A Narrative Review. Biomolecules 2022, 12, 1443. [Google Scholar] [CrossRef]
  156. Gaucher, F.; Bonnassie, S.; Rabah, H.; Marchand, P. Review: Adaptation of Beneficial Propionibacteria, Lactobacilli, and Bifidobacteria Improves Tolerance Toward Technological and Digestive Stresses. Front. Microbiol. 2019, 10, 841. [Google Scholar] [CrossRef] [PubMed]
  157. Megur, A.; Daliri, E.B.-M.; Baltriukienė, D.; Burokas, A. Prebiotics as a Tool for the Prevention and Treatment of Obesity and Diabetes: Classification and Ability to Modulate the Gut Microbiota. Int. J. Mol. Sci. 2022, 23, 6097. [Google Scholar] [CrossRef] [PubMed]
  158. Herbst, T.; Sichelstiel, A.; Schär, C.; Yadava, K.; Bürki, K.; Cahenzli, J.; McCoy, K.; Marsland, B.J.; Harris, N.L. Dysregulation of Allergic Airway Inflammation in the Absence of Microbial Colonization. Am. J. Respir. Crit. Care Med. 2011, 184, 198–205. [Google Scholar] [CrossRef]
  159. Abrahamsson, T.R.; Jakobsson, H.E.; Andersson, A.F.; Björkstén, B.; Engstrand, L.; Jenmalm, M.C. Low Diversity of the Gut Microbiota in Infants with Atopic Eczema. J. Allergy Clin. Immunol. 2012, 129, 434–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Nylund, L.; Satokari, R.; Nikkilä, J.; Rajilić-Stojanović, M.; Kalliomäki, M.; Isolauri, E.; Salminen, S.; de Vos, W.M. Microarray Analysis Reveals Marked Intestinal Microbiota Aberrancy in Infants Having Eczema Compared to Healthy Children in At-Risk for Atopic Disease. BMC Microbiol. 2013, 13, 12. [Google Scholar] [CrossRef] [Green Version]
  161. Bustamante, M.; Oomah, B.D.; Oliveira, W.P.; Burgos-díaz, C.; Rubilar, M.; Shene, C. Probiotics and Prebiotics Potential for the Care of Skin, Female Urogenital Tract, and Respiratory Tract. Folia Microbiol. 2020, 65, 245–264. [Google Scholar] [CrossRef] [Green Version]
  162. Nambiar, R.B.; Perumal, A.B.; Shittu, T.; Sadiku, E.R.; Sellamuthu, P.S. Editorial: Probiotics, Prebiotics, Synbiotics, Postbiotics, & Paraprobiotics—New Perspective for Functional Foods and Nutraceuticals. Front. Nutr. 2023, 10, 1164676. [Google Scholar] [CrossRef]
  163. Pu, J.; Hang, S.; Liu, M.; Chen, Z.; Xiong, J.; Li, Y.; Wu, H.; Zhao, X.; Liu, S.; Gu, Q.; et al. A Class IIb Bacteriocin Plantaricin NC8 Modulates Gut Microbiota of Different Enterotypes In Vitro. Front. Nutr. 2022, 9, 877948. [Google Scholar] [CrossRef]
  164. Sornsenee, P.; Chatatikun, M.; Mitsuwan, W.; Kongpol, K.; Kooltheat, N.; Sohbenalee, S.; Pruksaphanrat, S.; Mudpan, A.; Romyasamit, C. Lyophilized Cell-Free Supernatants of Lactobacillus Isolates Exhibited Antibiofilm, Antioxidant, and Reduces Nitric Oxide Activity in Lipopolysaccharide- Stimulated RAW 264.7 Cells. PeerJ 2021, 9, e12586. [Google Scholar] [CrossRef]
  165. Scarpellini, E.; Rinninella, E.; Basilico, M.; Colomier, E.; Rasetti, C.; Larussa, T.; Santori, P.; Abenavoli, L. From Pre-and Probiotics to Post-Biotics: A Narrative Review. Int. J. Env. Res. Public. Health 2022, 19, 37. [Google Scholar] [CrossRef]
  166. Raman, M.; Ambalam, P.; Kondepudi, K.K.; Pithva, S.; Kothari, C.; Patel, A.T.; Purama, R.K.; Dave, J.M.; Vyas, B.R.M. Potential of Probiotics, Prebiotics and Synbiotics for Management of Colorectal Cancer. Gut Microbes 2013, 4, 181–192. [Google Scholar] [CrossRef] [Green Version]
  167. Hong, L.; Lee, S.M.; Kim, W.S.; Choi, Y.J.; Oh, S.H.; Li, Y.L.; Choi, S.H.; Chung, D.H.; Jung, E.; Kang, S.K.; et al. Synbiotics Containing Nanoprebiotics: A Novel Therapeutic Strategy to Restore Gut Dysbiosis. Front. Microbiol. 2021, 12, 715241. [Google Scholar] [CrossRef]
  168. Bandyopadhyay, B.; Das, S.; Mitra, P.K.; Kundu, A.; Mandal, V.; Adhikary, R.; Mandal, V.; Mandal, N.C. Characterization of Two New Strains of Lactococcus Lactis for Their Probiotic Efficacy over Commercial Synbiotics Consortia. Braz. J. Microbiol. 2022, 53, 903–920. [Google Scholar] [CrossRef]
  169. Polakowski, C.B.; Kato, M.; Preti, V.B.; Schieferdecker, M.E.M.; Ligocki Campos, A.C. Impact of the Preoperative Use of Synbiotics in Colorectal Cancer Patients: A Prospective, Randomized, Double-Blind, Placebo-Controlled Study. Nutrition 2019, 58, 40–46. [Google Scholar] [CrossRef]
  170. Sergeev, I.N.; Aljutaily, T.; Walton, G.; Huarte, E. Effects of Synbiotic Supplement on Human Gut Microbiota, Body Composition and Weight Loss in Obesity. Nutrients 2020, 12, 222. [Google Scholar] [CrossRef] [Green Version]
  171. Salminen, S.; Collado, M.C.; Endo, A.; Hill, C.; Lebeer, S.; Quigley, E.M.M.; Sanders, M.E.; Shamir, R.; Swann, J.R.; Szajewska, H.; et al. The International Scientific Association of Probiotics and Prebiotics (ISAPP) Consensus Statement on the Definition and Scope of Postbiotics. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 649–667. [Google Scholar] [CrossRef]
  172. Thorakkattu, P.; Khanashyam, A.C.; Shah, K.; Babu, K.S.; Mundanat, A.S.; Deliephan, A.; Deokar, G.S.; Santivarangkna, C.; Nirmal, N.P. Postbiotics: Current Trends in Food and Pharmaceutical Industry. Foods 2022, 11, 3094. [Google Scholar] [CrossRef]
  173. Cuevas-González, P.F.; Liceaga, A.M.; Aguilar-Toalá, J.E. Postbiotics and Paraprobiotics: From Concepts to Applications. Food Res. Int. 2020, 136, 109502. [Google Scholar] [CrossRef]
  174. Overmyer, K.A.; Rhoads, T.W.; Merrill, A.E.; Ye, Z.; Westphall, M.S.; Acharya, A.; Shukla, S.K.; Coon, J.J. Proteomics, Lipidomics, Metabolomics, and 16S DNA Sequencing of Dental Plaque from Patients with Diabetes and Periodontal Disease. Mol. Cell. Proteom. 2021, 20, 100126. [Google Scholar] [CrossRef]
  175. Mayorgas, A.; Dotti, I.; Salas, A. Microbial Metabolites, Postbiotics, and Intestinal Epithelial Function. Mol. Nutr. Food Res. 2021, 65, 2000188. [Google Scholar] [CrossRef]
  176. Joshi, C.; Patel, P.; Kothari, V. Anti-Infective Potential of Hydroalcoholic Extract of Punica Granatum Peel against Gram-Negative Bacterial Pathogens. F1000Research 2019, 8, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Alessandri, G.; Milani, C.; Duranti, S.; Mancabelli, L.; Ranjanoro, T.; Modica, S.; Carnevali, L.; Statello, R.; Bottacini, F.; Turroni, F.; et al. Ability of Bifidobacteria to Metabolize Chitin-Glucan and Its Impact on the Gut Microbiota. Sci. Rep. 2019, 9, 5755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. He, Z.; Ma, Y.; Yang, S.; Zhang, S.; Liu, S.; Xiao, J.; Wang, Y.; Wang, W.; Yang, H.; Li, S.; et al. Gut Microbiota-Derived Ursodeoxycholic Acid from Neonatal Dairy Calves Improves Intestinal Homeostasis and Colitis to Attenuate Extended-Spectrum β-Lactamase-Producing Enteroaggregative Escherichia Coli Infection. Microbiome 2022, 10, 79. [Google Scholar] [CrossRef] [PubMed]
  179. Kao, H.J.; Balasubramaniam, A.; Chen, C.C.; Huang, C.M. Extracellular Electrons Transferred from Honey Probiotic Bacillus Circulans Inhibits Inflammatory Acne Vulgaris. Sci. Rep. 2022, 12, 19217. [Google Scholar] [CrossRef]
  180. Islam, M.I.; Seo, H.; Redwan, A.; Kim, S.; Lee, S.; Siddiquee, M.; Song, H.Y. In Vitro and In Vivo Anti-Clostridioides Difficile Effect of a Probiotic Bacillus Amyloliquefaciens Strain. J. Microbiol. Biotechnol. 2022, 32, 46–55. [Google Scholar] [CrossRef]
  181. Ishnaiwer, M.; Bezabih, Y.; Javaudin, F.; Sassi, M.; Bemer, P.; Batard, E.; Dion, M. In Vitro and In Vivo Activity of New Strains of Bacillus Subtilis against ESBL-Producing Escherichia Coli: An Experimental Study. J. Appl. Microbiol. 2022, 132, 2270–2279. [Google Scholar] [CrossRef]
  182. Cui, L.H.; Yan, C.G.; Li, H.S.; Kim, W.S.; Hong, L.; Kang, S.K.; Choi, Y.J.; Cho, C.S. A New Method of Producing a Natural Antibacterial Peptide by Encapsulated Probiotics Internalized with Inulin Nanoparticles as Prebiotics. J. Microbiol. Biotechnol. 2018, 28, 510–519. [Google Scholar] [CrossRef]
  183. Hashem, N.M.; Hosny, N.S.; El-Desoky, N.I.; Shehata, M.G. Effect of Nanoencapsulated Alginate-Synbiotic on Gut Microflora Balance, Immunity, and Growth Performance of Growing Rabbits. Polymers 2021, 13, 4191. [Google Scholar] [CrossRef]
  184. Jung, Y.J.; Kim, H.S.; Jaygal, G.; Cho, H.R.; Lee, K.B.; Song, I.B.; Kim, J.H.; Kwak, M.S.; Han, K.H.; Bae, M.J.; et al. Postbiotics Enhance NK Cell Activation in Stress-Induced Mice through Gut Microbiome Regulation. J. Microbiol. Biotechnol. 2022, 32, 612–620. [Google Scholar] [CrossRef]
  185. Golkar, N.; Ashoori, Y.; Heidari, R.; Omidifar, N.; Abootalebi, S.N.; Mohkam, M.; Gholami, A. A Novel Effective Formulation of Bioactive Compounds for Wound Healing: Preparation, In Vivo Characterization, and Comparison of Various Postbiotics Cold Creams in a Rat Model. Evid.-Based Complement. Altern. Med. 2021, 2021, 8577116. [Google Scholar] [CrossRef]
  186. Puccetti, M.; Gomes dos Reis, L.; Pariano, M.; Costantini, C.; Renga, G.; Ricci, M.; Traini, D.; Giovagnoli, S. Development and In Vitro-In Vivo Performances of an Inhalable Indole-3-Carboxaldehyde Dry Powder to Target Pulmonary Inflammation and Infection. Int. J. Pharm. 2021, 607, 121004. [Google Scholar] [CrossRef]
Figure 1. Conceptual hypothesis model concerning the rise of HGT mechanisms and the intrinsic resistance of pathogens during biofilm formation, which occurs during infections.
Figure 1. Conceptual hypothesis model concerning the rise of HGT mechanisms and the intrinsic resistance of pathogens during biofilm formation, which occurs during infections.
Futurepharmacol 03 00034 g001
Figure 2. Plant extract administration via oral, transdermal, and intravenous methods (https://mindthegraph.com/ accessed on 3 May 2023).
Figure 2. Plant extract administration via oral, transdermal, and intravenous methods (https://mindthegraph.com/ accessed on 3 May 2023).
Futurepharmacol 03 00034 g002
Figure 3. Chemical pathway of the formation of gluconic acid and hydrogen peroxide (H2O2) in honey.
Figure 3. Chemical pathway of the formation of gluconic acid and hydrogen peroxide (H2O2) in honey.
Futurepharmacol 03 00034 g003
Figure 4. Osmotic mechanism related to the antibacterial activity of honey. (A) Proposed mechanism of osmotic action mediated by honey sugars, as shown by the blue arrows. (B) S. aureus cells treated with honey showed cell collapse caused by dehydration, as indicated by the cells with the blue arrows (photos obtained by the authors of the present study).
Figure 4. Osmotic mechanism related to the antibacterial activity of honey. (A) Proposed mechanism of osmotic action mediated by honey sugars, as shown by the blue arrows. (B) S. aureus cells treated with honey showed cell collapse caused by dehydration, as indicated by the cells with the blue arrows (photos obtained by the authors of the present study).
Futurepharmacol 03 00034 g004
Table 1. Plant extract patents and their applications.
Table 1. Plant extract patents and their applications.
TitleDatePatentsCountryPlantsBacteriaApplicationReferences
Antibacterial essential oil2023CN115708794AChinaGrape, Zedoariae rhizoma, Radix angelicae pubescentis, myrrh, Ligusticum wallichii, Eucalyptus globulus, Boswellia carterii, clove, peppermint, and corianderStreptococcus pyogenes, S. aureus, and K. pneumoniaeIt inhibits the formation of a biofilm on the surface of a biological material.[53]
An antibacterial and anti-inflammatory composition containing plant extracts that provides itching relief, to be applied accordingly2022CN115844777AChinaBasil, bergamot, Salvia miltiorrhiza, witch hazel, aloe, mint, juniper berry, camellia seed, calendula, Polygonum multiflorum, honeysuckle, camphor tree, pseudo-ginseng, honeysuckle, olive, camellia, tea, daphne, Gentiana rigescens, Polygonatum kingianum, licorice, and ChrysanthemumEscherichia coli, S. aureus, and Candida albicansIt can be used in oral care products, medicines, and skin care products, it provides relief from itching, and it has antibacterial and anti-inflammatory effects.[54]
Plant composite antibacterial agent containing peony extracts, to be prepared and applied accordingly2022CN115399343BChinaScutellaria baicalensis, aloe, selfheal, honeysuckle, Pogostemon cablin, oregano, clove, lavender, and Folium artemisiae argyi or Forsythia suspensaS. aureus, C. albicans, E. coli, and P. AeruginosaIt is highly sanitary and safe, and it has low metal corrosion and low skin irritation.It has fast-acting, highly efficient, and long-lasting properties.[55]
Antibacterial hand sanitizer composition containing plant extracts2021KR102424044B1South KoreaLeek, green onion, purslane, water parsley, and perilla leavesE. coli, S. aureus, and S. epidermidisIt has excellent moisturizing abilities to help maintain skin health. Moreover, it is possible to formulate a hand sanitizer with excellent sterilization power.[56]
Plant antibacterial mite-killing agent, to be prepared accordingly, and it can be used as a daily essential 2021CN112868678AChinaThymus vulgaris, rosemary, Sophora flavescens, Folium artemisiae argyi, licorice, and dandelion E. coli, S. aureus, and C. albicansThe active molecules of the plant extract can act on the brain nerve cells of the mites, thus stimulating the brain neurons, and enabling the mites to enter a deep sleep; this will achieve the effect of efficiently killing the mites.[57]
Composition for improving antibacterial, anti-inflammatory, antiviral, and immune functions, comprising the extract of ligularia stenocephala as an active ingredient2020WO2021182661A1South KoreaLigularia stenocephalaE. coli, P. aeruginosa, Aspergillus niger, Staphylococcus hominis, Bacillus subtilis, and Streptococcus pneumoniaeIt may be offered as a health-functional food composition or a pharmaceutical composition that enhances antibacterial, anti-inflammatory, and immune functions.[58]
Plant extract compositions and methods to make and use plant extract compositions2020US20210386074A1United StatesGinger, green coffee, rosemary, and honeysuckleKlebsiella aerogenes, and S. aureusThe extract composition of the present invention may have general or broad-spectrum disinfectant efficacy.[59]
Plant extract hydrolysates and an antibacterial product containing plant extract hydrolysates2009US9138451B2United StatesEquiseti, Juglandis, Millefolii, Quercus, Taraxaci, Althaeae, Matricariae, Centaurium, Levisticum, Rosmarinus, Angelica(e), Artemisia, Astragalus, Leonurus, Salvia, Saposhnikovia, Scutellaria, Siegesbeckia, Armoracia, Capsicum, Cistus, Echinacea, Echinacea, Galphimia, and HederaS. aureus, S. epidermidis, S. pyogenes,
S. pneumoniae, Streptococcus mutans, and Haemophilus influenzae
It can be used to
produce agents with antibacterial effects against severe infections.
[60]
Antibacterial composition comprising plant extracts2002WO2003035093A1South KoreaFoeniculum vulgare, Illicium verum, Asarum heterotropoides, CinnamomumCandida and
Trichophyton sp.
An antifungal composition that is safe for skin and has superior antifungal activity. It can be applied as a cleaner, a treating agent for dermatomycosis, such as athlete’s foot, a disinfectant, among other uses. [61]
Table 2. Geographical and floral origin of honey, as well as the main pathogens and methods used in the study of the antimicrobial activity of honey.
Table 2. Geographical and floral origin of honey, as well as the main pathogens and methods used in the study of the antimicrobial activity of honey.
Geographical OriginFloral OriginBacterial StrainAnalytical MethodReferences
AustraliaE. marginata, E. patens, E. platypus, E. wandoo
Banksia spp., Callistemon spp., Corymbia calophylla, Leptospermum subtenue, and Leptospermum scoparium
S. aureus ATCC 25923, ATCC 29213 and NCTC 10442, S. epidermidis ATCC 11047, E. faecalis ATCC 29212 and ATCC 51299, A. baumannii ATCC 7844, E. coli ATTC 25922, P. aeruginosa ATCC, 27853, and Salmonella enterica subsp. enterica serovar Typhimurium ATCC 13311Kirby–Bauer Test
Minimal inhibitory concentration (MIC) and minimal bactericidal concentration (MBC)
[85]
E. macrorrhynchaAlcaligenes faecalis, Citrobacter freundii, E. coli, K.
aerogenes, K. pneumoniae, Mycobacterium phlei, Salmonella enterica subsp. enterica serovar California, Salmonella enterica subsp. enterica serovar Enteritidis, S. Typhimurium, Shigella sonnei, S. aureus and S. epidermidis, Serratia marcescens, and C. albicans
MIC
(agar dilution method)
[86]
ColombiaPolyfloral honeyE. coli ATCC 25922, S. Typhimurium ATCC 14028, and Listeria monocytogenes ATCC 19118Agar diffusion technique[87]
CubaTurbina corymbosa (L.) Raf, Ipomoea triloba L., Avicennia germinans Jacq., Govania polygama (Jack) Urb, and Lysiloma latisiquum (L.) BenthB. subtilis ATCC 6633 and S. aureus
ATCC 25923
Minimum active dilution (MAD) via the agar incorporation technique[88]
Polyfloral honeyClinical isolates: S. aureus 13, S. epidermidis 35, S. pneumoniae 9, S. pyogenes 12, S. pyogenes C-105, S. pyogenes m46, Streptococcus agalactiae 1357, Streptococcus mitis 22, Streptococcus oralis 1235, Streptococcus anginosus 2513, Streptococcus parasanguinis 2761, Streptococcus salivarius 14, Streptococcus gordonii 143, E. faecalis 212, E. faecium 17, L. monocytogenes 49, Enterobacter cloacae 19902, C. freundii 55, Salmonella enterica subsp. enterica serovar Fyris 3813, S. marcescens 28315, A. baumanii 8, K. pneumoniae 15, P. aeruginosa 24. E. coli 23, Proteus mirabilis 112, and C. albicans 18MAD via the agar incorporation technique[89]
Polyfloral honeyClinical isolates: S. aureus 13, S. epidermidis 35, S. pneumoniae 9, S. pyogenes 12, S. pyogenes C-105, S. pyogenes m46, S. agalactiae 1357, S. mitis 22, S. oralis 1235, S. anginosus 2513, S. parasanguinis 2761, S. salivarius 14, S. gordonii 143, E. faecalis 212, E. faecium 17, L. monocytogenes 49, E. cloacae 19902, C. freundii 55, S. Fyris 3813, S. marcescens 28315, A. baumanii 8, K. pneumoniae 15, P. aeruginosa 24. E. coli 23, P. mirabilis 112, and C. albicans 18MAD via the agar incorporation technique
Inhibition of biofilm formation and removal of preformed biofilm assay
Transmission Electron Microscopy (TEM) for morphology analysis
[90]
EcuadorEucalyptus spp.S. aureus ATCC 25923, S. pyogenes ATCC 19615, P. aeruginosa ATCC 27853, E. coli ATCC 25922, and C. albicans ATTC 90028MAD via the agar incorporation technique[91]
S. aureus CAMP and K. pneumoniae KPC 609803Inhibition of biofilm formation and removal of preformed biofilm assays[80]
Eucalyptus spp.S. aureus MRSA ATCC 2592, P. aeruginosa ATCC 2785, and S. aureus MRSA S21 (clinical isolate)Inhibition of biofilm formation and removal of preformed biofilm assay
TEM for morphology analysis
[92]
Eucalyptus spp.S. aureus, MRSA S21 (clinical isolate), and P. aeruginosa P28 clinical isolate)Inhibition of biofilm formation and removal of preformed biofilm assay
TEM for morphology analysis
[79]
Persea americanaS. aureus CAMP and K. pneumoniae KPC 609803Inhibition of biofilm formation and removal of preformed biofilm assay[80]
Brassica napusS. aureus CAMP and K. pneumoniae KPC 609803Inhibition of biofilm formation and removal of preformed biofilm assay[80]
IranEucalyptus spp.E. coli ATCC 25922, P. aeruginosa ATCC 27853, S. aureus ATCC 25923, and E. faecalis ATCC 11700Kirby–Bauer Test[93]
ItalyEucalyptus spp.S. aureus subsp. aureus ATCC 9144, and P. aeruginosa ATCC 27853Kirby–Bauer Test[94]
KenyaPolyfloral honeyClinical isolates: S. aureus 13, S. epidermidis 35, S. pneumoniae 9, S. pyogenes 12, S. pyogenes C-105, S. pyogenes m46, S. agalactiae 1357, S. mitis 22, S. oralis 1235, S. anginosus 2513, S. parasanguinis 2761, S. salivarius 14, S. gordonii 143, E. faecalis 212, E. faecium 17, L. monocytogenes 49, E. cloacae 19902, C. freundii 55, S. Fyris 3813, S. marcescens 28315, A. baumanii 8, K. pneumoniae 15, P. aeruginosa 24, E. coli 23, P. mirabilis 112, and C. albicans 18MAD via the agar incorporation technique
Inhibition of biofilm formation and removal of preformed biofilm assay
TEM for morphology analysis
[90]
MauritiusEucalyptus spp.E. coli (clinical isolate), E. coli ATCC 25922, Proteus spp. (clinical isolate), P. mirabilis ATCC 12453, Pseudomonas spp. (clinical isolate), P. aeruginosa ATCC 27853, Klebsiella spp. (clinical isolate), Streptococcus spp. (clinical isolate), and S. epidermidis ATCC 35984 and ATCC 14990Kirby–Bauer Test[95]
New ZealandLeptospermum scopariumClinical isolates: S. aureus 13, S. epidermidis 35, S. pneumoniae 9, S. pyogenes 12, S. pyogenes C-105, S. pyogenes m46, S. agalactiae 1357, S. mitis 22, S. oralis 1235, S. anginosus 2513, S. parasanguinis 2761, S. salivarius 14, S. gordonii 143, E. faecalis 212, E. faecium 17, L. monocytogenes 49, E. cloacae 19902, C. freundii 55, S. Fyris 3813, S. marcescens 28315, A. baumanii 8, K. pneumoniae 15, P. aeruginosa 24. E. coli 23, P. mirabilis 112, and C. albicans 18MAD via the agar incorporation technique[89]
Leptospermum scoparium var. Incanum, Leptospermum scoparium var. incanum + Kunzea ericoides
Leptospermum scoparium var. incanum + Kunzea ericoides, and Trifolium spp.
S. aureus NCTC 8325 and ATCC 25923, S. aureus HA-MRSA, and S. aureus CA-MRSAInhibition of biofilm formation and removal of preformed biofilm assays[96]
PakistanZiziphus mauritiana, Azadirachta
indica, Ziziphus spina-christi, Citrus sinensis, and Brassica
nigra
Salmonella enterica subsp. enterica serovar TyphiMBC, MIC, and agar well diffusion assays[97]
PolandPrunus spinosa L., Polyfloral honey, Salix spp., Brassica napus L., Phacelia tanacetifolia Benth., Solidago vigaurea L., and Helianthus spp.E. coli D31 (CGSC 5165), Bacillus circulans ATCC 61; S. aureus, 1-KI (clinical isolate), P. aeruginosa (ATCC 27853), P. aeruginosa 02/18 (clinical isolate), A. niger 71, Saccharomyces cerevisiae, and C. albicansMAD via the agar incorporation technique[98]
SlovakiaCrataegus laevigata,
Abies alba Mill, and
Robinia pseudoacacia
P. mirabilis and E. cloacaeMAD via the agar incorporation technique
Removal of preformed biofilm
assay
[99]
Robinia pseudoacacia, Rubus spp., Brassica napus, Rubus idaeus, and Phacelia spp. P. aeruginosa CCM1960 and S. aureus CCM4223MIC and MBC assays[100]
SpainEucalyptus spp.S. aureus and MRSA (clinical isolate), S. pyogenes, E. coli, and P. aeruginosa (clinical isolate)Disk–plate diffusion method[101]
TurkeyNigella sativa L.E. coli ATCC 25,922, E. faecalis ATCC 29,121, S. aureus ATCC 6538, S. enteric subsp. enterica ATCC 14,028/363–154, B. subtilis B209, Bacillus cereus, and L. monocytogenes ATCC 7677Disc diffusion and MIC assays[102]
Table 4. Summary of recent in vivo and in vitro studies reporting beneficial effects for the host and the antimicrobial activities of prebiotics, probiotics, synbiotics, and postbiotics.
Table 4. Summary of recent in vivo and in vitro studies reporting beneficial effects for the host and the antimicrobial activities of prebiotics, probiotics, synbiotics, and postbiotics.
Type of BioticsCompounds and/or SpeciesAntimicrobial ActivityReferences
PrebioticsPunica granatum peel extractIn vitro and in vivo Caenorhabditis elegans nematode model assays demonstrated a reduction in hemolytic activity and biofilm formation caused by P. aeruginosa and they promoted the growth of B. bifidum and L. plantarum probiotic strains.[176]
Chitin-glucan (CG)In vitro assays showed high levels of growth in all bifidobacterial species, particularly the Bifidobacterium breve 2L isolate in the in vivo Groningen rat model, which became more abundant in the gut of B. breve 2L.[177]
Phthalyl pullulan nanoparticles (PPNs)A gut dysbiosis-induced murine model was used, and their restorative effect in the eubiosis microbiota was assessed using the pathogen, E. coli K99. [167]
Ursodeoxycholic acid (UDCA)In vitro assays demonstrated a reduction in E. coli serotype O101:H9 growth, proinflammatory effects in Caco-2 cells, and cell integrity damage. Moreover, in vivo assays used on neonatal mice model also exhibited attenuated colitis symptoms and recovered colonic short-chain fatty acid (SCFA) production.[178]
ProbioticsTwo Lactococcus lactis subsp. lactis strainsIn vivo bacterial feeding of these probiotic strains for 30 days in a Swiss albino mice model was conducted, and they improved gut colonization and IgA levels.[168]
Honey probiotic B. circulans isolateIn vitro co-culture of B. circulans and Cutibacterium acnes significantly suppressed pathogen growth. Moreover, in vivo assays using the ICR mice model, B. circulans, generated electrons that inhibited C. acnes growth and diminished inflammation.[179]
Bacillus amyloliquefaciens (BA PMC-80)In vitro co-culture assays of BA PMC-80 and Clostridioides difficile demonstrated significant pathogen inhibition; an in vivo hamster model exhibited no toxicity, a less severe infection, and late death. [180]
Two new strains of B. subtilis (CH311 and S3B)In vitro gut model demonstrated the ability of B. subtilis CH311 and S3B to reduce ESBL-E. coli titers using 4 log CFU/mL; however, the in vivo murine model showed no reduction in the ESBL-E. coli fecal titers. [181]
SynbioticsPediococcus acidilactici plus phthalyl inulin nanoparticles (PINs)In vitro antimicrobial activity was tested using a cocultivation assay. A statistical reduction of more than 3 log CFU/mL of Salmonella enterica subsp. enterica serovar Gallinarum, together with P. acidilactici plus PINs, was observed when compared with the control and the PINs or probiotic groups alone.[182]
Encapsulated S. cerevisiae plus Moringa oleifera leaf extract (MOLE)In vivo rabbit model revealed no effects on interleukin-l or IgG and IgA levels, and it showed a significantly higher number of beneficial microbes. Moreover, a significant increase in in vitro inhibitory activities was observed against E. coli BA 12296B, S. aureus NCTC 10788, C. albicans ATCC MYA-2876, L. monocytogenes ATCC 19116, and Salmonella enterica subsp. enterica serovar Senftenberg ATCC 8400.[183]
PPNs plus L. plantarumA gut dysbiosis-induced murine model was used and the E. coli K99 infection was markedly suppressed after several well-known beneficial bacteria, including Lactobacillus and Bifidobacterium, were incrementally introduced.[167]
Two L. lactis subsp. lactis strains with inulinThe Swiss albino mice model exhibited a significant reduction in IgA levels that is comparable with commercial probiotics and prebiotic consortiums on the market.[168]
PostbioticsLyophilized cell-free supernatants (LCFS) of Lactobacillus isolatesDemonstrated strong inhibition and eradication antibiofilm activities for A. baumannii and E. coli, and a reduction in nitric oxide production in the RAW 264.7 cell line was observed.[164]
L. plantarum KM1, L. plantarum KM2, Bacillus velezensis KMU01 postbiotics mixtures 1:1:1 (vol/vol)NK cell activation was significantly higher in the C57BL/6N mice model, and TNF-α levels in the RAW264.7 cell line was significantly reduced when compared with the LPS.[184]
Individual LCFS of L. fermentum, L. reuteri, and B. subtilis sp. natto in a postbiotic cold creamAll postbiotic cold creams exhibited different degrees of immunomodulatory, anti-inflammatory, and antimicrobial activities in the Sprague Dawley rat model when compared with controls (no treatment and only cold cream).[185]
Indole-3-carboxaldehyde (3-IAld) (a microbial tryptophan metabolite)The use of 3-IAld inhalable dry powder demonstrated optimal pulmonary administration and toxicological safety, also reducing aspergillosis scores by acting on the infection and inflammation sites.[186]
Legend—CFU: Colony-forming units; IgA: Immunoglobulin A; ICR: Crl:CD1; Vol/vol: Volume/volume; NK: Natural killer; TNF-α: Tumor necrosis factor alpha; LPS: Lipopolysaccharides.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Machado, A.; Zamora-Mendoza, L.; Alexis, F.; Álvarez-Suarez, J.M. Use of Plant Extracts, Bee-Derived Products, and Probiotic-Related Applications to Fight Multidrug-Resistant Pathogens in the Post-Antibiotic Era. Future Pharmacol. 2023, 3, 535-567. https://doi.org/10.3390/futurepharmacol3030034

AMA Style

Machado A, Zamora-Mendoza L, Alexis F, Álvarez-Suarez JM. Use of Plant Extracts, Bee-Derived Products, and Probiotic-Related Applications to Fight Multidrug-Resistant Pathogens in the Post-Antibiotic Era. Future Pharmacology. 2023; 3(3):535-567. https://doi.org/10.3390/futurepharmacol3030034

Chicago/Turabian Style

Machado, António, Lizbeth Zamora-Mendoza, Frank Alexis, and José Miguel Álvarez-Suarez. 2023. "Use of Plant Extracts, Bee-Derived Products, and Probiotic-Related Applications to Fight Multidrug-Resistant Pathogens in the Post-Antibiotic Era" Future Pharmacology 3, no. 3: 535-567. https://doi.org/10.3390/futurepharmacol3030034

Article Metrics

Back to TopTop