Next Article in Journal
Vegan Ice Cream Made from Soy Extract, Soy Kefir and Jaboticaba Peel: Antioxidant Capacity and Sensory Profile
Next Article in Special Issue
Functional Fermented Milk with Fruit Pulp Modulates the In Vitro Intestinal Microbiota
Previous Article in Journal
Analysis of Volatile Compounds, Composition, and Thermal Behavior of Coffee Beans According to Variety and Roasting Intensity
Previous Article in Special Issue
Probiotic Oxalate-Degrading Bacteria: New Insight of Environmental Variables and Expression of the oxc and frc Genes on Oxalate Degradation Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Use of Bacteriocins and Bacteriocinogenic Beneficial Organisms in Food Products: Benefits, Challenges, Concerns

by
Svetoslav Dimitrov Todorov
1,*,
Igor Popov
2,
Richard Weeks
3 and
Michael Leonidas Chikindas
2,3,4
1
ProBacLab, Laboratório de Microbiologia de Alimentos, Departamento de Alimentos e Nutrição Experimental, Faculdade de Ciências Farmacêuticas, Universidade de São Paulo, São Paulo 05508-000, SP, Brazil
2
Center for Agrobiotechnology, Don State Technical University, 344002 Rostov-on-Don, Russia
3
Health Promoting Naturals Laboratory, School of Environmental and Biological Sciences, Rutgers State University, New Brunswick, NJ 08904, USA
4
Department of General Hygiene, I.M. Sechenov First Moscow State Medical University, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Foods 2022, 11(19), 3145; https://doi.org/10.3390/foods11193145
Submission received: 18 September 2022 / Revised: 6 October 2022 / Accepted: 7 October 2022 / Published: 10 October 2022

Abstract

:
This review’s objective was to critically revisit various research approaches for studies on the application of beneficial organisms and bacteriocins as effective biopreservatives in the food industry. There are a substantial number of research papers reporting newly isolated bacterial strains from fermented food products and their application as potential probiotics, including partial characterization of bacteriocins produced by these microorganisms. Most of these studies follow scientific community-accepted standard procedures and propose various applications of the studied strains and bacteriocins as potential biopreservatives for the food industry. A few investigations go somewhat further, performing model studies, exploring the application of expressed bacteriocins in a designed food product, or trying to evaluate the effectiveness of the studied potential probiotics and bacteriocins against foodborne pathogens. Some authors propose applications of bacteriocin producers as starter cultures and are exploring in situ bacteriocin production to aid in the effective control of foodborne pathogens. However, few studies have evaluated the possible adverse effects of bacteriocins, such as toxicity. This comes from well-documented reports on bacteriocins being mostly non-immunogenic and having low cytotoxicity because most of these proteinaceous molecules are small peptides. However, some studies have reported on bacteriocins with noticeable cytotoxicity, which may become even more pronounced in genetically engineered or modified bacteriocins. Moreover, their cytotoxicity can be very specific and is dependent on the concentration of the bacteriocin and the nature of the targeted cell. This will be discussed in detail in the present review.

1. Introduction

The search for optimal manufacturing conditions (raw materials and their processing, storage, distribution of the final product, etc.) has always been an objective of the food industry. Producing food commodities at an optimal price, with extended shelf life, covering consumer demand for healthier ingredients, and without chemical preservatives can be supported by using antimicrobials of microbial origin, such as bacteriocins. Bacteriocins are easy to produce, generally safe, and specific in their mode of action, qualities that should make them perfect antimicrobials for the food industry. However, are bacteriocins really an answer to the demand for safe, nontoxic, and effective antimicrobial preservatives for the food industry? Since the discovery of bacteriocins, the search for bacteriocins with potential applications in the food industry has been the subject of more than 13276 papers (www.scopus.com, accessed 29 August 2022). European regulatory authorities authorized the use of nisin in 1983 [1] for application in some food products as an effective biopreservative for the control of spoilage organisms. Five years later, the FDA approved the use of nisin in the USA (FDA 21CFR184.1538; https://www.accessdata.fda.gov/scripts/cdrh/cfdocs/cfcfr/cfrsearch.cfm?fr=184.1538, accessed on 25 August 2022). Worldwide, nisin’s approved applications cover a wide variety of products with often very different permitted levels of use (for more information, see Todorov, Franco, Tagg [2]). However, out of the many bacteriocins that have been investigated, only nisin and later pediocin PA-1 (for reviews, see Silva et al. [3] and Soltani et al. [4]) are approved for use in the food industry as partially purified bacteriocins. What is the reason for such limited use, with only two bacteriocins selected from a large cohort of potent antimicrobial bacteriocins? Is it because many (if not most) isolated and characterized bacteriocins are not suited for large-scale production and application? Or, perhaps, because other bacteriocins are less potent than those already utilized by the food industry? Are there strong lobbying and industry pressures behind the production of nisin? Or is there more scrutiny in approving new food additives so that novel bacteriocins can no longer meet regulatory requirements? Whatever the answer is, research on the isolation, characterization, and proposal of new bacteriocins as potentially effective biopreservatives is a growing scientific trend in need of common guidelines and a consensus on appropriate methods and reporting standards (Figure 1).
Moreover, in addition to nisin and pediocin PA-1, some other bacteriocins were applied as part of multicomponent commercial preparations, recently reviewed by Chikindas et al. [5], most of them associated with the application of nisin A (BioSateTM, Chr. Hansen) produced by Lactococcus lactis subsp. lactic BS-10, applied for the control of different cheeses and the prevention of off-flavors and late blowing associated with clostridia contaminations; undefined bacteriocins (HOLDBACKTM, Dupont), produced by Propionibacterium freudenreichii subsp. shermanii DSM 706 and Lacteicaseibacillus rhamnosus DSM 7061, for the inhibition of mold and psychrotrophs in cottage cheeses; Sakacin A and pediocin PA-1/AcH (BactofermTM F-LC, Chr. Hansen), produced by Latilactobacillus curvatus, Pediococcus acidilactici, and Staphylococcus xylosus strains, for the control of Listeria monocytogenes and as a meat fermentation starter; plantaricin and carnocin (ALCMix1, Danisco DuPont), produced by Lactiplantibacillus plantarum and Staphylococcus carnosus, for the control of listerial contaminants in fermented sausages and cooked ham; leucocin (BactofermTM B-SF-43, Chr. Hansen), produced by Leuconostoc carnosum, for the control of listeria in vacuum and modified atmosphere stored meat products; and sakacin (BactofermTM B-2 and BactofermTM B-FM, Chr. Hansen), produced by Latilactobacillus sakei (alone or in combination with Staphylococcus xylosus), for the control of listeria in stored or fresh meat products [5].
Therefore, we aimed to be critical in discussing different research approaches for performing studies on the application of bacteriocinogenic lactic acid bacteria (LAB) and bacteriocins as effective biopreservatives for the food industry.

2. Bacteriocins: What Good Do They Do for Us in Food Products?

Production of antimicrobial metabolites, including bacteriocins, is an evolutionary response resulting in the creation of effective survival mechanisms so microbial species can compete, communicate, and protect themselves in multimicrobial environments [6]. Bacteriocins are “bacterially produced peptides that are active against other bacteria and against which the producer has a specific immunity mechanism” [7]. The functional role of bacteriocins in natural environments (soil, water reservoirs, etc.) is not yet fully understood. We often refer to their antimicrobial properties as their key feature, playing a role in protecting ecological niches and competing with other “neighboring” microorganisms (for a recent in-depth review, see Heilbronner et al. [8]). However, the signaling role of bacteriocins and their involvement in quorum sensing processes may be their primary role in the life of a microbial community [5].
Bacteriocins from LAB are peptides (small proteins) with specific chemical and physical characteristics that influence their specific interactions with targeted cells [9]. Generally, bacteriocins are active against organisms closely related to the producing species [2]. However, a few examples from the last two decades provide evidence that some bacteriocins produced by LAB can be inhibitory against specific Gram-negative bacterial species, yeasts and fungi [2], Mycobacterium [10,11], and even some viruses [12,13]; for a review with emphasis on a proposed strategy for the use of bacteriocinogenic probiotics, see Tiwari et al. [14]. Such reports are an interesting contribution to our knowledge of bacteriocins; however, it must be noted that, in several cases, these nonspecific activities may be a consequence of the combined action between bacteriocins and some elements of the culturing environment [15]. Even if this is the case, the combined application of bacteriocins and additional synergetic factors against clinically relevant pathogens and spoilage microorganisms merits scientific attention and the further development of effective therapeutic and biopreservative agents (see reviews Cavera et al. [16], Chikindas et al. [5], and Arthur et al. [17]).
The successful application of bacteriocins in the food industry as a factor for food safety can be associated with the extension of shelf life and protection against foodborne pathogens, especially in nonrefrigerated products. Consequently, there is a significantly reduced risk of transmission of pathogens and spoilage microorganisms from these products. In addition to reducing the number of foodborne pathogens in food products, bacteriocins significantly reduce economic losses associated with food spoilage, outbreaks, and recalls. It is important to mention that bacteriocins are defined as natural antimicrobial agents and, thus, fit into the consumers’ growing requests for fewer chemical preservatives and treatments and a growing trend toward more natural food products [4].
Nisin is one of the most extensively studied bacteriocins and is used in a wide variety of food products as an effective factor in the inhibition and even elimination of foodborne pathogens and spoilage microorganisms in dairy, meat, fish, seafood, fruits, vegetables, cereals, and their derivatives [18,19,20,21,22,23,24].
When applied as biopreservatives in the food industry, bacteriocins (alone and in combination with other natural preservatives, such as bacteriophages) are shown to effectively reduce the presence of spoilage microorganisms and foodborne pathogens [25]. For example, the industrial application of bacteriocins and bacteriocinogenic strains is a clear proof that they can contribute to the reduction of Clostridium spp. in dairy products and pasteurized eggs [20,26]; Listeria monocytogenes in fish, dairy, and meat products [27]; and vancomycin-resistant enterococci in different food products [28]. A clear advantage of bacteriocins compared with other antimicrobials and conventional sanitization, conservation, and sterilization processes is their selectivity and antimicrobial spectrum of activity [2]. Bacteriocins normally possess a rather narrow spectrum of activity and are carefully selected for a specific purpose. They can inhibit selected spoilage and foodborne pathogens, while microorganisms associated with technological purposes, such as starter cultures, remain unaffected [29]. In contrast, most of the sanitizing, conservation, and sterilization processes kill practically all vegetative cells and some spore forms of any microorganisms present in the food product, including technologically important starter, adjunct, or probiotic cultures.

3. Effect of Environmental Factors on the Bacteriocin Effectiveness

Interactions between bacteriocins and other components of the food matrix are a topic that has been investigated by some researchers but still needs much more attention. A research approach focused on bacteriocin–food matrix interactions will give the ability to predict the efficacy of studied bacteriocins in a specific food matrix. It is known that most bacteriocins are cationic molecules and are reported as being amphipathic [5,7]. Based on their chemical specificity, bacteriocins can interact with target organisms. Similarly, they can also have specific behaviors in the food environment. Components of the food system can inhibit or promote bacteriocin activity, and these potential interactions need to be evaluated for any new bacteriocins and their application in specific food systems. A significant number of papers report on the high efficacy of nisin in dairy products [30,31,32], and the industrial application of nisin in the biopreservation of different soft cheeses, dairy spreads, and other dairy products has been explored [33,34,35]. However, nisin is less effective in meat products [34,36,37,38]. It has been suggested that the interaction of nisin with lipids from meat products can “trap” the bacteriocin and reduce its efficacy. However, other bacteriocins are effective in a meat environment. For instance, Lactiplantibacillus plantarum (formerly Lactobacillus plantarum) 423, producer of plantaricin 423, a class IIa bacteriocin, was studied as a promising biopreservative culture in the preparation of different types of salami [39]. Results showed that L. monocytogenes, used as a test organism (sensitive to plantaricin 423), was significantly reduced to the level of no detection when Lb. plantarum 423 was used as a starter–adjuvant culture in the preparation of salami. However, the application of a non-bacteriocinogenic mutant of Lb. plantarum 423 did not affect the growth of L. monocytogenes [39]. Numerous authors have evaluated the application of bacteriocins or bacteriocin-producing strains in meat preservation, and these works suggest a beneficial effect of bacteriocins as antimicrobials on the reduction of foodborne pathogens in the final products [37,38,40,41,42,43,44,45,46,47,48,49].

4. Bacteriocins: Safety Is the Priority

Another reasonable question is related to the appropriate choice of the cell line used for evaluating the cytotoxicity of bacteriocins. Most studies use Caco-2 as a model cell line for determining the cytotoxicity of different molecules, including bacteriocins. However, are cancer cell lines really a good model? Vero cells were utilized in several studies on the cytotoxicity of different bacteriocins [12,13,33,50,51,52,53], and the human hepatocellular carcinoma cell line, Huh7.5, has also been used [53]. The pharmaceutical industry recommends specific cell lines to determine the cytotoxicity of different pharmaceutical molecules, and the determination of bacteriocin cytotoxicity should adhere to the same or at least equivalent standards.
The term ‘cytotoxicity’ generally refers to a broad and ill-defined meaning in the drug discovery and pharmaceutical development industry. In the case of in vitro cell culture assays, for a compound or treatment to be considered cytotoxic, there needs to be evidence that it can interfere with cellular attachment, significantly alter morphology, adversely affect cell growth rate, or cause cell death [54,55]. Various assays have been developed and used for measuring viability or cytotoxicity in vitro, including classical dye inclusion or exclusion and colony formation assays [55,56,57,58]. Despite the variety of assays available for determining cytotoxicity not only for bacteriocins but different pharmaceutical and cosmetic products, well-accepted standards are fact and have been established [55,59].
While bacteriocins are generally considered to be nontoxic for mammalian cells, some examples easily poke holes in this assumption. Cytolysin is an antimicrobial peptide produced by several strains of Enterococcus spp. and showed toxicity at high concentrations [60]. Cytotoxicity needs to be considered as a milestone test when the safety of bacteriocins is being evaluated prior to their use in the food industry or in human or veterinary medicine. However, specific cytotoxicity against targeted cells can also be an advantage and should be explored for designing drugs targeting specific mammalian cells, such as cancer cells [61,62,63].
There was no reported evidence of acute toxicity when nisin was orally administered at 1 g/kg in rats [64]. However, Vaucher et al. [65] reported on possible toxicity in mice as evidenced by histological changes in the spleen, skin, and liver after exposure to 0.825 mg/kg nisin for 21 days. Moreover, the authors described that these histological changes were associated with specific, possibly inflammatory processes that it will be difficult to differentiate between the effect of purified nisin and the high salt content of commercial nisin (Nisaplin®) used in the mentioned study [65]. Extrapolating this to humans and considering that the weight of an average man is approximately 70 kg, this would be equivalent to a daily uptake of 50–70 g of nisin, an amount that is much higher than the realistic potential intake of nisin from different food products where nisin was applied as a biopreservative. Sahoo et al. [66] reported that the LD50 of the bacteriocin TSU4, which is produced by Ligilactobacillus (formerly Lactobacillus) animalis TSU4, is higher than 200 mg/kg when orally administered to mice. Moreover, after oral administration for 21 days at 0.5 mg/kg, the bacteriocin TSU4 showed neither mortality in mice nor significant changes in physiological conditions and immunological markers, an argument for the absence of a toxic effect for the studied bacteriocin [66].
Additional arguments for the safety of pediocin N6, another bacteriocin, were reported by Marlida et al. [67], who administered up to 20,000 mg/kg of pediocin N6 in mice. Lactocin 160 is a bacteriocin produced by Lacticaseibacillus (formerly Lactobacillus) rhamnosus 160 with activity against different pathogens associated with bacterial vaginosis. Lactocin 160 was intravaginally applied in rabbits at 18 mg per treatment, and no evidence of acute toxicity was observed. Furthermore, in vitro and in vivo safety evaluation experiments showed that lactocin 160 is not associated with any severe irritation of vaginal epithelial tissue [68]. A bacteriocin produced by Enterococcus faecalis, enterocin AS-48, showed no toxicity when intraperitoneally administered to mice (100 g/mouse, corresponding to 5 mg/kg) every 8 h over a 2-day period, and no skin sensitization or allergic contact dermatitis was observed [69].
However, studies that would seem to show the opposite of the above-mentioned results have also been reported: nisin and pediocin PA-1 were cytotoxic when applied at higher concentrations against Vero cell lines [70,71]. Moreover, it was reported that a semi-purified bacteriocin produced by Lb. plantarum ST8SH was highly cytotoxic at a concentration of 25 μg/mL, but not at 5 μg/mL [72].
Since bacteriocins are polypeptides (small-sized proteins), it is assumed that proteolytic enzymes in the stomach and other parts of the intestinal tract can hydrolyze them, rendering them inert. However, what about the upper regions of the GIT? What about the protective effect from various components of food matrixes? All these are subjects that merit further attention and deeper analyses.
The sensitivity of bacteriocins to the effects of proteolytic enzymes present in the GIT of humans and other animals can be discussed in terms of both positive and negative characteristics. On one side, when applied in the biopreservation of food products, leftover residues of bacteriocins can be digested by proteolytic enzymes in the GIT, lowering the risk of cytotoxicity. However, if the target of bacteriocins is pathogens within the GIT, several barriers may be involved and affect their stability and biological activity. It is well-established in different in vitro and in vivo studies that orally administered bacteriocins and even those produced in situ are inactivated or digested by proteolytic enzymes naturally present in the GIT [73,74,75]. Kheadr et al. [76] reported that pediocin PA-1 was stable in the stomach. However, it was degraded when exposed to the conditions of the small intestine. It has been suggested that class I bacteriocins may be more resistant to proteases compared with class II bacteriocins due to undergoing extensive post-translational modifications [77,78]. It was reported that nisin can remain partly stable when exposed to gastrointestinal fluids; however, bioengineered bacteriocins or encapsulation can improve stability [79]. Nisin bioengineering of the C-terminal region can improve the bacteriocin’s activity and stability in the presence of proteolytic enzymes [80,81]. Since most bacteriocins are expected to perform at the end of the small intestine and in the colon, encapsulation was suggested as a potential approach for the effective protection and controlled delivery of these molecules [17,82,83,84].
The immune response to the administration of bacteriocins is considered an essential part of evaluating the safety of antimicrobial peptides and other food additives or pharmaceutical preparations [85,86]. Pediocin AcH (also known as pediocin PA-1) [87] and TSU4 [66] were shown to be non-immunogenic. However, when administrated over more extended periods of time, a nisin-containing preparation (that also contains milk proteins and salt) can be associated with a significant increase in macrophages and monocytes isolated from the peripheral blood [88]. After the administration of pyocin S5, low levels of antibodies were recorded in the serum of the test animals [89,90].
As part of the safety evaluation of bacteriocins, it is important to observe no adverse effects on the integrity of tight junctions or cell-to-cell adhesion [91]. Different cell lines mimicking the GIT or the intestinal mucosa have been utilized in the evaluation of the effect of bacteriocins on tight junctions, including HCT-8, HT29, and Caco-2 [92,93,94,95]. According to Belguesmia et al. [91], nisin and enterocin S37 can be considered safe since no significant reduction in epithelial integrity was observed when both bacteriocins were supplemented at 10 μg/mL to Caco-2/TC7 cells. Moreover, when Caco-2 cells were treated with 2 μg/mL of divercin AS7 for 24 h, no changes in the integrity of the cells were observed [96]. In a similar experiment, the interaction between Caco-2 cells and plantaricin A showed that the bacteriocin can protect the cells’ integrity and tight junctions [94].

5. Interactions between Bacteriocins and Environmental Factors

The role of environmental factors (temperature, pH, presence of chemicals) has been the subject of several studies related to bacteriocins produced by various LAB [44,97,98,99,100,101]. In addition to evaluating optimal conditions for bacteriocin’s action, these reasonably simple methods can predict the efficacy of new bacteriocins and substantially reduce research costs and time. On the other side, the specificity of a bacteriocin’s adsorption to and effect on target pathogens is very useful information to know which foodborne pathogens and spoilage organisms can be targeted during the biopreservation processes.
Bacteriocin adsorption to target cells can be based on the recognition of specific receptors and charge-based interactions or interference with hydrophobic surface molecules and is considered the first step of the antibacterial mode of action (for a review, see, for instance, Drider et al. [102], and for a study of model interactions, see Soliman et al. [103]). To penetrate the cell membrane and disturb cellular integrity, some, if not all, bacteriocins need to be effective in their recognition of specific receptors or/and to have strong physicochemical interactions with target cells. This process may include a specific target such as lipid II, recognized by lantibiotics, and some antimicrobial proteins of eukaryotic origin (for review, see Grein et al. [104]). However, a lack of interaction with lipid II is reported for some lantibiotics as well; moreover, different groups of lantibiotics have distinct interactions with lipid II as a target (comprehensively reviewed by Wang et al. [105]). Other targets (recognition sites, “anchoring molecules”) were reported for different bacteriocins, such as mannose phosphotransferase for pediocin PA-1, a class IIa bacteriocin [106]. These complex processes are influenced by several factors, including pH, temperature, and the presence of chemical agents. All of these factors can influence physiological conditions and the integrity of the cell membrane and, consequently, can influence the interaction of bacteriocins with target receptors or their direct contact with the cell membrane.
The in-depth evaluation of the influence of environmental factors on the adsorption of bacteriocins to target cells can be a valuable tool in predicting the efficacy of bacteriocins, not only in food biopreservation processes but also in the primary selection of potential therapeutically active antimicrobial agents for human and veterinary medicine application.

6. Application of Bacteriocins

How can we apply bacteriocin(s) in food biopreservation processes? There is consumer demand for better food products: products with a longer shelf life, without chemical preservatives, more natural ingredients, etc. Food retail businesses provide choices to the consumers and offer food products that fit into the concept of “bio products” or “organic products”. However, the final choice of consumers is often driven by the price of the product. In addition, the safety of “bioproducts” or “organic products” is debatable and has already been the subject of discussion and evaluation [107,108]. If applying natural biopreservatives, including and focusing on bacteriocins, increases the final product’s price, will this food product be competitive on the retail market? This is why most research groups agree that applying highly purified bacteriocins is not an economically feasible option. A more realistic approach is to apply partially purified bacteriocins or guarantee in situ bacteriocin production by applying bacteriocin producers as starter or adjunct cultures in the fermentation process (Table 1).
The perfect scenario then is to develop one culture that will be responsible for solving all the needs from the food technology point of view (such as to be an excellent starter culture) that, at the same time, can be a producer of antimicrobial metabolites and will guarantee food safety and microbial quality by inhibiting potential spoilage and foodborne pathogens. This will be a perfect scenario, but it is unrealistic and unlikely to be achievable, almost the same as the desire of alchemists to find a Panacea, a universal drug for all illnesses. Starter cultures and biopreservative cultures have different roles in fermentation processes. Most of the LAB starter cultures play a role in the acidification of their environments [128,129] and produce a variety of enzymes (including proteolytic and lipolytic enzymes), particularly important in the preparation of fermented meat products [130,131]. In dairy products, proteolytic enzymes produced by LAB starter cultures can be actively involved in the reduction of allergenicity by hydrolysis of specific proteins, particularly caseins [132,133]. There is ongoing research on the use of LAB starter cultures’ enzymatic systems to reduce gluten in fermented cereal products [134]. Bacteriocins can be important players in food safety. However, they need to be produced at reasonable levels, especially considering the temperature during the preparation of food products, availability of nutrients for the bioprotective LAB culture, etc. They must exhibit their activity without being affected by proteolytic enzymes produced by the starter cultures [34].
There are several other issues that are worth mentioning. It is important in cases where the bacteriocinogenic culture will be applied as an adjunct starter culture that the specific culture does not influence the organoleptic characteristics of the final product while also being safe for application in food and feed products. Safety measures are essential and need to include physiological and biochemical tests for antibiotic susceptibility, hemolytic activity, and potential production of biogenic amines [135,136]. The presence of virulence-related genetic determinants should be addressed as well. Concerns about the possible transfer of those virulence genes have been the subject of several studies and models. It should be mentioned that Suvorov [137] reported on the low probability of possible horizontal gene transfer between enterococci and other LAB or pathogens in the GIT. However, safety issues related to the presence of virulence factors in the starter and biopreservative cultures need to be evaluated as well, bearing in mind that specific microbial cultures will be exposed to very different conditions, and it is the researcher’s responsibility to guarantee that none of them are carrying virulence determinants in any known forms.
Cereal-based fermented products are popular and consumed all around the world. The food industry has been adapting existing traditional products or developing new cereal-based food products (fermented or not) while taking into consideration consumer preferences. The Balkan Peninsula is known for one of its traditional fermented cereal-based beverages, boza [138]. This product is traced back to the ancient times and is highly appreciated by different ethnic groups because of its nutritional value and as a part of traditional medicine in the region [138,139]. Boza has been described as a source of numerous bacteriocin producers [138,139,140,141,142,143]. Specific antimicrobial characteristics for some of the boza-derived LAB and their bacteriocins have even been linked to antiviral, anti-Mycobacterium tuberculosis, and anti-fungal activity [45,144]. Interactions between bacteriocins produced by LAB, part of the boza starter culture consortium, and spoilage microorganisms or foodborne pathogens were evaluated, and the role of antimicrobial proteins was highlighted [46,145].
To evaluate the potential of two different bacteriocinogenic strains (E. faecium ST88 and Enterococcus mundtii CRL35) as potential biopreservative cultures, Pingitore et al. [146] performed a model study of cheese preparation and applied both strains as adjunct cultures. In a series of preliminary tests, the authors have shown the behavior of the studied bacteriocinogenic strains in laboratory conditions relating to their interactions with L. monocytogenes (the target foodborne pathogen in this project). In addition, the behavior of bacteriocins and interactions with the target organism (L. monocytogenes) were evaluated at different temperatures and pH, and in the presence of dairy ingredients and commonly used additives to better understand and predict effectiveness in upcoming in situ experiments. Pingitore et al. [146] clearly showed that the level of bacteriocin production (6400 AU/mL for E. faecium ST88 and 25600 AU/mL for E. mundtii CRL35), recorded against L. monocytogenes 426, can play a critical role in controlling the targeted foodborne pathogen. We need to acknowledge that the level of bacteriocin production in a commercial MRS medium and a dairy environment (in milk) cannot be the same, and this caveat has been discussed in several papers [146,147,148].
Effective antimicrobial action against different strains of the target pathogen is necessary to ensure that the proposed bacteriocin can protect the food product. The selected bacteriocinogenic strain and bacteriocin should be highly effective against targeted pathogens but, at the same time, should not inhibit the growth of the beneficial microbiota. This has been discussed and pointed out by several research papers. Pingitore et al. [146] performed experiments on the adsorption of bacteriocins (produced by E. faecium ST88 and E. mundtii CRL35) to L. monocytogenes under different conditions, including various temperatures, pH, and in the presence of cheese-making additives, and the E. mundtii CRL35 bacteriocin was shown to be superior as compared with the activity of the bacteriocin from E. faecium ST88. We need to keep in mind that for any bacteriocin to act as an antimicrobial, the first step will be to adhere to the surface of the target strain. This may depend on the recognition of a specific receptor or can be a nonspecific adherence to the cell wall [7]. For that specific bacteriocin produced by E. mundtii CRL35 [146], the technological conditions of the cheese-making procedure, including conditions such as low pH and production, maturation, and refrigeration temperatures, were optimal for bacteriocin adsorption to L. monocytogenes. This study presents possible positive results for the control of L. monocytogenes by bacteriocins produced by E. faecium ST88 and E. mundtii CRL35 with potentially better performance by the bacteriocin mundticin (produced by E. mundtii CRL35).

7. How Effective Can the Application of Partially Purified Bacteriocins Be?

Nisin is a great example of the successful application of bacteriocins (lantibiotic, class I, classification by Heng et al. [149]). It has been nearly one century since the discovery of nisin, and the long history of its safe food-associated applications can be attributed to its extremely low cytotoxicity, similar to that of NaCl [150], and its sensitivity to digestive proteases and absence of any influence on the sensory properties of food [151].
Commercial applications of nisin are regulated by the European Union Food Safety Authorities, where it is licensed as a food preservative (E234). The FAO/WHO Codex Committee on milk and milk products has authorized nisin to be applied as a food additive in processed cheeses, with a limit of 12.5 mg (as calculated for pure nisin) per kilogram of the product [152]. However, commercial nisin is a partially purified industrial preparation. The commercial preparation of nisin follows the strict rules and regulations developed by international and national authorities following the recommendations of the FAO and WHO. When we apply semi-purified commercial bacteriocin preparations, attention must be given to the composition of the supporting components and ballast ingredients.
Moreover, Vaucher et al. [65] pointed out the fact that some histological changes after application of Nisaplin® can lead to specific inflammatory processes, most probably not due to the effect of nisin itself but from the high salt content in the industrial formulation (Nisaplin®) [65]. In commercial preservation processes, different salts and milk residues in the Nisaplin® formulation can be responsible for the observed side effects. In addition, the application of nisin is a practice in the biopreservation of pasteurized eggs, canned food, some meat, and bakery products [30,32]. However, when applied to food products, bacteriocin(s) and bacteriocinogenic formulation ingredients (such as bacteriocinogenic starter cultures, protective cultures, etc.) need to be clearly stated on the label to alert the consumers to the potential presence of allergenic ingredients (dairy or salts) in the product.

8. Limitations of Use of Bacteriocin Producers

The application of bacteriocin producers as adjuncts or starter cultures can be considered a perfect scenario for biopreservation processes, which has been proposed by several research papers suggesting that specific LAB can provide technological (for performing biotransformation of raw products to a fermented food product) properties, in addition to biopreservation properties, producing different bioactive molecules and contributing to the food safety of the final product [34,72,131,133]. However, relatively few have been implemented in practical industrial processes. The principal problem may be an absence of communication or collaboration between academia and industry. Scientific problems related to applying LAB as live cultures with biopreservative properties can also be related to the fact that some bacteriocin producers cannot meet the criteria for safety assessments. The appropriate regulatory governmental organization controls the applications of starter and adjunct cultures to guarantee the safety of food products regionally and internationally. The WHO, EFSA, and FDA are some examples of these agencies.
Bacteriocin production is reported in different LAB [7]. However, not all LAB are considered safe. A good example is found in species belonging to the genera Streptococcus, where the GRAS status is granted only for Streptococcus thermophilus. The rest of the Streptococcus spp. are not considered safe, and some are serious human and animal pathogens and are even related to some types of cancers, as reported for some Streptococcus bovis strains [153,154,155]. Enterococcaceae is yet another good example. Some species are well-known and have been used for centuries as starter cultures, especially in the Mediterranean region, in the production of different types of cheeses and salami [156]. On the other hand, numerous Enterococcus spp. can be directly linked to nosocomial infections, and vancomycin-resistant Enterococcus is a serious health problem [157]. Some infections were directly linked to medical cases related to endocarditis, bacteremia, and intra-abdominal, pelvic, urinary tract, and central nervous system infections [158]. The presence of vancomycin-resistant enterococci, most probably related to the use of avoparcin as a feed additive in animal husbandry, is a significant threat to public health since glycopeptides are considered to be the last line of drugs in the treatment of enterococcal infections [159,160,161,162,163]. Antibiotic resistance alone cannot explain the virulence of enterococci. The process is more complex, and most infections follow a common sequence of events involving colonization and adhesion to host tissues, invasion of the tissue, and resistance to both nonspecific and specific defense mechanisms of the host. The pathogen must produce pathological changes either directly by toxin production or indirectly by inflammation [164].
Studies of the presence of virulence factors and antibiotic resistance genes have been conducted for different LAB [136,165,166] and showed that some “GRAS species” (although the GRAS status is usually assigned to a strain) can carry some of these genes, compromising their safety. The general question, then, is whether the bacteriocin producer can be defined as safe. Comprehensive safety evaluations need to be performed for any strain before application in a specific part of the food production process to guarantee the safety of the consumers.
Most LAB are commonly considered safe based on their long history of use in different fermented food processes [167]. However, some LAB are reported as pathogens (some reports are Land et al. [168], Kulkarni and Khoury [169], Salminen et al. [170], Karime et al. [171], and an old but comprehensive review: Aguirre and Collins [172]) and food spoilage microorganisms (for instance, Kalschne et al. [173] and Andreevskaya et al. [174]).
A different problem is related to a more practical issue. When working with bacteriocin production, most of the studies have been performed using a very rich growth medium such as MRS [97,99,100,148]. Several studies were reported on the optimization of bacteriocin production, applying simple approaches such as exclusion or replacement of some media components, or building sophisticated mathematical models and following the effect of different constituents of the growth medium and cultivation parameters on bacteriocin production [97,99,100,148]. Different byproducts have been proposed as alternatives for the low-cost production of bacteriocins of varying types and origins [97,148]. The question, however, remains: will LAB be able to produce sufficient levels of bacteriocins and at the specific desirable amount to effectively protect against food spoilage? Will bacteriocin be able to reach the entire food system to guarantee its efficacy? Studies have shown that some of these problems can be considered solved, as the authors reported on the efficacy of bacteriocins and their production in situ [175]. However, we cannot always expect high bacteriocin expression levels across different food environments. The concentration and availability of nutritional components can be limited in addition to competition with other microorganisms. The problem, however, is that many studies are not adequately planned and executed, as there is an obvious conflict between optimal conditions for bacterial growth, the production of bacteriocins, and the conditions needed when performing model studies. For instance, many studies were conducted under refrigeration conditions. Under these conditions, L. monocytogenes can multiply, but this is far below the temperature limits for LAB, and we cannot expect LAB to produce bacteriocins if they experience low-temperature stress.
In summary, when planning to apply specific bacteriocinogenic LAB as live adjunct cultures, we need to be sure that the fermented food product will provide specific nutrients related to bacterial growth and antimicrobial protein production. Additionally, the technological temperature of the production of food products must be in accordance with the specific temperature needs of the bacteriocin-producing LAB, neither very low (as refrigeration) nor very high (as part of cooking or sterilization of the product).

9. Spectrum of Activity: Kill Only Bad and Not Good

One of the biggest advantages of bacteriocins as antimicrobial agents is their relatively high selectivity [5]. Antibiotics are defined as chemotherapeutic agents with activity against microorganisms (bacteria and fungi) and protozoa [176]. Antibiotics are highly effective in controlling pathogenic bacteria. However, they often kill the beneficial microbiota as well. Cases of antibiotic-related diarrhea as a consequence of the treatment of clinical cases in patients are a serious problem in hospitals. Clinical practices recommend the application of vitamins from group B, supporting beneficial microbiota treatment, enforcement with selected probiotics, and, in some cases, the consumption of traditional fermented food products rich in LAB [177].
The application of antibiotics is strictly controlled in animal feed as growth promoters and is authorized only as therapeutic agents for the treatment of sick animals and not as prophylactic treatment in veterinary practice. The uncontrolled application of antibiotics in veterinary practice in the past is one of the well-recognized reasons for the rise of antibiotic resistance. According to the European Commission in 2005, more than 10 million tons of antibiotics have been released into the biosphere since the start of commercial production. This has exerted a very strong selection on the development of resistant strains. Resistance may be inherent to a bacterial genus or species but may also be acquired through the exchange of genetic material, mutations, and the incorporation of new genes [178]. Teuber et al. [179] suggested that starter cultures and probiotics may serve as vectors in the transfer of antibiotic resistance genes. Such a transfer has been documented in other bacterial groups by Levy and Marshall [180].
The specific influence of bacteriocins on the GIT microbiota needs to be considered as an important issue, since long-term application may have a negative effect on diversity and consequently on function and may even be associated with different diseases, including inflammatory immune diseases, functional bowel disorders, insulin resistance, and obesity [181]. The effects of bacteriocins on changes in the diversity of intestinal microbiota in different regions of the GIT were reported by Kheadr et al. [76], Le Blay et al. [182], Guinane et al. [183], and Le Lay et al. [184] by different in vitro models. Treatment with nisin and pediocin PA-1 did not result in significant changes in the commensal GIT microbiota for in vitro evaluations [182,183,184] and in vivo model studies [185,186]. On the other hand, the opposite was reported in an in vitro study [187], where lacticin 3147 (class I bacteriocin) induced a shift in the microbiota from Firmicutes to Proteobacteria.
Umu et al. [175] have discussed the production of bacteriocins as part of the arsenal of the probiotic potential of LAB in preventing the growth of pathogens in gut environments. The authors evaluated the potential of five bacteriocin-producing strains of LAB (producers of sakacin A; pediocin PA-1; enterocins P, Q, and L50; plantaricins EF and JK; and garvicin ML) and their isogenic nonproducing mutants for probiotic potential in an animal model via 16S rRNA gene analysis. When combined with appropriate pre-selection analysis based on an evaluation of pH, temperature, and environmental conditions, this can be the ideal approach to making the right choice when selecting potentially effective bacteriocinogenic strains.
In an animal model study on rats, Bernbom et al. [185] reported on the use of Lc. lactis (nisin producer) or purified nisin that increased the presence of bifidobacteria and decreased the number of enterococci in fecal samples. However, in a different study on bacterial cultures associated with human microbiota, nisin A and nisin Z showed inhibitory effects on bifidobacteria and lactobacilli. When the effect of pediocin was evaluated, no changes in the two mentioned microorganisms’ presence were observed [188]. Similar results were reported by Dabour et al. [186] when they evaluated the effect of a bacteriocin on bifidobacteria and lactobacilli in a model animal study in mice. Moreover, Dobson et al. [189] reported that oral administration of Lc. lactis DPC6520, a lacticin 3147 producer, did not affect diversity and proportions within the microbiota of the distal colon. Similar arguments for the effect of bacteriocin Abp118, produced by Ligilactobacillus (formerly Lactobacillus) salivarius UCC118As, were presented by Riboulet-Bisson et al. [190], showing no significant influence on the proportion of the general microbial communities of mouse gut microbiota, with a significant decrease only in Spirochetes and Firmicutes. In a similar experiment using a mouse model, Murphy et al. [191] reported that when applied, some specific bacteriocinogenic strains did not significantly impact Firmicutes but may increase the presence of Bacteroidetes and Proteobacteria and reduce representatives of Actinobacteria.
In a murine animal model, administration of bacteriocinogenic Lb. salivarius UCC118 showed an increase in Peptococcaceae and a reduction in the proportion of Riknelleaceae and Porphyromonadaceae [192]. Previous examples are evidence that the bacteriocinogenic strains and bacteriocins themselves can impact the functional diversity of the colonic microbiota. However, the impacts of bacteriocins and bacteriocin-producing strains are not yet fully understood, and additional in vitro and in vivo studies need to be designed, including the application of different strains and bacteriocins together with different animal models.
Even if several studies demonstrated the beneficial role of bacteriocins in the control of spoilage microorganisms and foodborne pathogens, pointing to an important role of the narrow spectrum of activity of bacteriocins and their effectiveness against targeted species, some authors have raised their concerns about the application of bacteriocins associated with the fact that these antimicrobials may be too narrow in their spectrum of activity [23]. Moreover, bacteriocins can be trapped by lipids due to their hydrophobic nature, which are naturally present in many food products [193]. From a general point of view, predicting what specific spoilage and/or foodborne pathogen should be targeted and which bacteriocin will have to be applied will be difficult.
Mills et al. [193] mentioned that the distribution of bacteriocins into the food matrix can be yet another limiting factor diminishing their effectiveness. However, targeting specific areas where spoilage of foodborne pathogens can be present in high numbers should be considered as an additional strategy for the application of bacteriocins as effective biopreservatives. Zhou et al. [194] suggested that plantaricin BM1 can be spread on the surface of hams and, in this way, can be more effective in inhibiting L. monocytogenes. Incorporating bacteriocins into the packaging material, from which they can be released and effectively inhibit targeted spoilage or pathogens, is yet another approach presently under investigation [195].
Although some may see it as a potential disadvantage, most authors agree that bacteriocins have the advantage of being rather selective in their spectrum of activity and being able to inhibit particular species. Selectivity is an advantage as most bacteriocins can be applied in an appropriate system for the control of specific targeted pathogens while leaving other parts of the microbial population unhindered. Selectivity is a key feature of antimicrobial peptides that cannot be associated with antibiotics or chemical antimicrobials. Udompijitkul et al. [196] suggested the application of nisin for the control of germinated spores of Clostridium perfringens in a rich nutrient environment. Bartoloni et al. [197] evaluated the effect of nisin against clinical isolates of Clostridium difficile in comparison with vancomycin and metronidazole by minimum inhibitory concentration (MIC), minimum bactericidal concentration (MBC), and time–kill studies. According to Bartoloni et al. [197], nisin was more active than the other agents, with an MIC90 of 0.256 mg/L and strong bactericidal activity. They suggest that nisin may be a promising agent for the management of C. difficile-associated diarrhea. Application of bacteriocins for the control of Clostridium spp. without influencing the beneficial microbiota is just one example of a beneficial application of antimicrobial proteins from LAB.

10. Conclusions

Throughout the human history, food has always played an essential role. Traditions, religion, society, and family have always been and will continue to relate to or revolve around food. For centuries, preparing and preserving food commodities has been an empiric practice, transmitted from one generation to the next. The life sciences, established in the 18th and 19th centuries, have generated a collection of fundamental and applied knowledge over time. From a 21st-century perspective, however, the preparation of food products is a complex process, based on the involvement of biochemistry, microbiology, physics, chemistry, engineering, etc. Over the last century, the food preservation processes became better understood, and the role of different antimicrobial metabolites started to be clarified. Nowadays, consumers are not only satisfied by the nutritional values of food products but also pay significant attention to the safety and health-promoting properties of food commodities (Figure 1). In essence, food needs to be affordable, nutritional, safe, and health-promoting. The role of bacteriocins from LAB, as part of the larger family of antimicrobial peptides, has been proven by their contribution to the biopreservation processes. Bacteriocins can be clearly associated with controlling spoilage and foodborne pathogens, extending shelf life, and replacing chemical additives and preservatives. The more we extend our knowledge of bacteriocinogenic LAB and their role in fermentation and biopreservation processes, the more we discover that additional scientific questions will need to be addressed and answers provided. Currently, a solidly built scientific structure stands behind the application of bacteriocins in food and other areas, but this structure needs to be maintained and built upon in order to provide consumers with safe, nutritional, and beneficial food in the future.
Driven by scientific curiosity and demand from the food and pharmaceutical industries and both consumers and patients, research on antimicrobial proteins, including bacteriocins, is facing new challenges regarding an affordable cost of production and the partial purification of antimicrobial proteins to concentrations that are appropriate for specific applications. Additional post-production modifications can be an option for increasing the stability of bacteriocins, improving delivery to target sites in the case of both infections and the biopreservation of foods. Bioengineering of existing bacteriocins and replacing specific amino acids or regions is one possible way of modulating or increasing their spectrum of activity and extending their application areas. The abovementioned scientific challenges in the study of bacteriocins are representative of some of the challenges that still need to be addressed through academic and applied research of antimicrobial peptides.

Author Contributions

Conceptualization, S.D.T. and M.L.C.; resources, S.D.T.; writing—original draft preparation, S.D.T.; writing—review and editing, S.D.T., M.L.C., and R.W.; language corrections: R.W.; visualization, I.P. All authors have read and agreed to the published version of the manuscript.

Funding

SDT was funded by the University of Sao Paulo, SP, Brazil. IP and MLC were supported, in part, by the Ministry of Science and Higher Education of the Russian Federation (Project Number 075-15-2019-1880). RW received no funds for this study.

Data Availability Statement

Not applicable.

Acknowledgments

To Mia Miau for her unconditional support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. EFSA Panel on Food Additives and Nutrient Sources added to Food (ANS); Younes, M.; Aggett, P.; Aguilar, F.; Crebelli, R.; Dusemund, B.; Filipic, M.; Frutos, M.J.; Galtier, P.; Gundert-Remy, U.; et al. Scientific Opinion on the safety of nisin (E 234) as a food additive in the light of new toxicological data and the proposed extension of use. EFSA J. 2017, 15, 5063. [Google Scholar] [CrossRef]
  2. Todorov, S.D.; Franco, B.D.G.M.; Tagg, J.R. Bacteriocins of Gram-positive bacteria having activity spectra extending beyond closely-related species. Beneficial Microbs. 2019, 10, 315–328. [Google Scholar] [CrossRef] [PubMed]
  3. Silva, C.C.G.; Silva, S.P.M.; Ribeiro, S.C. Application of bacteriocins and protective cultures in dairy food preservation. Front. Microbiol. 2018, 9, 594. [Google Scholar] [CrossRef]
  4. Soltani, S.; Hammami, R.; Cotter, P.D.; Rebuffat, S.; Said, L.B.; Gaudreau, H.; Bédard, F.; Biron, E.; Drider, D.; Fliss, I. Bacteriocins as a new generation of antimicrobials: Toxicity aspects and regulations. FEMS Microbiol. Rev. 2021, 45, fuaa039. [Google Scholar] [CrossRef]
  5. Chikindas, M.L.; Weeks, R.; Drider, D.; Chistyakov, V.A.; Dicks, L.M. Functions and emerging applications of bacteriocins. Curr. Opin. Biotechnol. 2018, 49, 23–28. [Google Scholar] [CrossRef] [PubMed]
  6. Cornforth, D.M.; Foster, K.R. Competition sensing: The social side of bacterial stress responses. Nat. Rev. Microbiol. 2013, 11, 285–293. [Google Scholar] [CrossRef]
  7. Cotter, P.D.; Hill, C.; Ross, R.P. Bacteriocins: Developing innate immunity for food. Nat. Rev. Microbiol. 2005, 3, 777–788. [Google Scholar] [CrossRef]
  8. Heilbronner, S.; Krismer, B.; Brötz-Oesterhelt, H.; Peschel, A. The microbiome-shaping roles of bacteriocins. Nat. Rev. Microbiol. 2021, 19, 726–739. [Google Scholar] [CrossRef]
  9. Alvarez-Sieiro, P.; Montalbán-López, M.; Mu, D.; Kuipers, O.P. Bacteriocins of lactic acid bacteria: Extending the family. Appl. Microbiol. Biotechnol. 2016, 100, 2939–2951. [Google Scholar] [CrossRef] [Green Version]
  10. Montville, T.J.; Chung, H.-J.; Chikindas, M.L.; Chen, Y. Nisin A depletes intracellular ATP and acts in bactericidal manner against Mycobacterium smegmatis. Lett. Appl. Microbiol. 1999, 28, 189–193. [Google Scholar] [CrossRef]
  11. Chung, H.-J.; Montville, T.J.; Chikindas, M.L. Nisin depletes ATP and proton motive force in mycobacteria. Lett. Appl. Microbiol. 2008, 31, 416–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Torres, N.I.; Noll, K.S.; Xu, S.; Li, J.; Huang, Q.; Sinko, P.J.; Wachsman, M.B.; Chikindas, M.L. Safety, formulation, and in vitro antiviral activity of the antimicrobial peptide subtilosin against herpes simplex virus type 1. Prob. Antimicrob. Prot. 2013, 5, 26–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Quintana, V.M.; Torres, N.I.; Wachsman, M.B.; Sinko, P.J.; Castilla, V.; Chikindas, M. Antiherpes simplex virus type 2 activity of the antimicrobial peptide subtilosin. J. Appl. Microbiol. 2014, 117, 1253–1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Tiwari, S.K.; Dicks, L.M.T.; Popov, I.V.; Karaseva, A.; Ermakov, A.M.; Suvorov, A.; Tagg, J.R.; Weeks, R.; Chikindas, M.L. Probiotics at war against viruses: What is missing from the picture? Front. Microbiol. 2020, 11, 1877. [Google Scholar] [CrossRef]
  15. Bingol, E.B.; Akkaya, E.; Hampikyan, H.; Cetin, O.; Colak, H. Effect of nisin-EDTA combinations and modified atmosphere packaging on the survival of Salmonella enteritidis in Turkish type meatballs. Cyta J. Food. 2018, 16, 1030–1036. [Google Scholar] [CrossRef] [Green Version]
  16. Cavera, V.L.; Arthur, T.D.; Kashtanov, D.; Chikindas, M.L. Bacteriocins and their position in the next wave of conventional antibiotics. Int. J. Antimicrob. Agents 2015, 46, 494–501. [Google Scholar] [CrossRef] [PubMed]
  17. Arthur, T.D.; Cavera, V.L.; Chikindas, M.L. On bacteriocin delivery systems and potential applications. Future Microbiol. 2014, 9, 235–248. [Google Scholar] [CrossRef]
  18. Delves-Broughton, J. Nisin and its application as a food preservative. Int. J. Dairy Technol. 1990, 43, 73–76. [Google Scholar] [CrossRef]
  19. Verma, A.; Banerjee, R.; Dwivedi, H.; Juneja, V.K. Bacteriocins: Potential in food preservation. In Encyclopedia of Food Microbiology; Academic Press: Cambridge, MA, USA; Elsivier Science: Amsterdam, The Netherlands, 2014; pp. 180–186. [Google Scholar] [CrossRef]
  20. Arqués, J.L.; Rodríguez, E.; Langa, S.; Landete, J.M.; Medina, M. Antimicrobial activity of lactic acid bacteria in dairy products and gut: Effect on pathogens. Bio.Med. Res. Int. 2015, 2015, 584183. [Google Scholar] [CrossRef]
  21. Gálvez, A.; Abriouel, H.; Lucas, R.; Grande Burgos, M.J. Bacteriocins for Bioprotection of Foods; CABI Books; CABI International: Wallingford, UK, 2011. [Google Scholar] [CrossRef]
  22. Gálvez, A.; Abriouel, H.; Omar, N.B.; Lucas, R. Food applications and regulation. In Prokaryotic Antimicrobial Peptides; Springer: New York, NY, USA, 2011; pp. 353–390. [Google Scholar]
  23. Gharsallaoui, A.; Oulahal, N.; Joly, C.; Degraeve, P. Nisin as a food preservative: Part 1: Physicochemical properties, antimicrobial activity, and main uses. Crit. Rev. Food Sci. Nutr. 2016, 56, 1262–1274. [Google Scholar] [CrossRef]
  24. Xavier, J.; Gopalan, N.; Ramana, K. Immobilization of lactic acid bacteria and application of bacteriocin for preservation of fruit juices and bacteriocin production. Defence Life Sci. J. 2017, 2, 231–238. [Google Scholar] [CrossRef] [Green Version]
  25. Rendueles, C.; Duarte, A.C.; Escobedo, S.; Fernández, L.; Rodríguez, A.; García, P.; Martínez, B. Combined use of bacteriocins and bacteriophages as food biopreservatives. A review. Int. J. Food Microbiol. 2022, 368, 109611. [Google Scholar] [CrossRef] [PubMed]
  26. Le Lay, C.; Dridi, L.; Bergeron, M.G.; Ouellette, M.; Fliss, I. Nisin is an effective inhibitor of Clostridium difficeile vegetative cells and spore germination. J. Med. Microbiol. 2016, 65, 169–175. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, H.; Hoover, D.G. Bacteriocins and their Food applications. Compr. Rev. Food Sci. Food Safety 2006, 2, 82–100. [Google Scholar] [CrossRef]
  28. Bucheli, J.E.V.; Fugaban, J.I.I.; Holzapfel, W.H.; Todorov, S.D. Combined action of antibiotics and bacteriocins against vancomycin-resistant enterococci. Microorganisms. 2022, 10, 1423. [Google Scholar] [CrossRef]
  29. Simons, A.; Alhanout, K.; Duval, R.E. Bacteriocins, antimicrobial peptides from bacterial origin: Overview of their biology and their impact against multidrug-resistant bacteria. Microorganisms 2020, 8, 639. [Google Scholar] [CrossRef]
  30. Ibarra-Sanchez, L.A.; El-Haddad, N.; Mahmoud, D.; Miller, M.J.; Karam, L. Advances in nisin use for preservation of dairy products. J. Dairy Sci. 2020, 103, 2041–2052. [Google Scholar] [CrossRef]
  31. Abdulhussain, K.R.; Razavi, S.H. Plantaricin bacteriocins: As safe alternative antimicrobial peptides in food preservation. J. Food Safety 2020, 40, e12735. [Google Scholar] [CrossRef]
  32. Oros-Florez, Z.S.; Garcia-Almendarez, B.E.; Barboza-Corona, J.E.; Salcedo-Hernandez, R. A fast micromethod for the estimation of nisin activity in a soft cheese. Int. J. Dairy Technol. 2019, 72, 282–286. [Google Scholar] [CrossRef]
  33. Martinez, R.R.C.; Wachsman, M.; Torres, N.I.; LeBlanc, J.G.; Todorov, S.D.; Franco, B.D.G.M. Biochemical, antimicrobial and molecular characterization of a noncytotoxic bacteriocin produced by Lactobacillus plantarum ST71KS. Food Microbiol. 2013, 34, 376–381. [Google Scholar] [CrossRef]
  34. Favaro, L.; Penna, A.L.B.; Todorov, S.D. Bacteriocinogenic LAB from cheeses—Application in biopreservation? Trends Food Sci. Technol. 2015, 41, 37–48. [Google Scholar] [CrossRef]
  35. Todorov, S.D. What bacteriocinogenic lactic acid bacteria do in the milk? In Raw Milk. Balance between Hazards and Benefits; Chapter 8; Nero, L.A., de Carvalho, A.F., Eds.; Elsevier: Amsterdam, The Netherlands; Academic Press: Cambridge, MA, USA, 2018; pp. 149–174. [Google Scholar] [CrossRef]
  36. Barbosa, M.S.; Todorov, S.D.; Ivanova, I.V.; Belguesmia, Y.; Choiset, Y.; Rabesona, H.; Chobert, J.-M.; Haertlé, T.; Franco, B.D.G.M. Characterization of a two-peptide plantaricin produced by Lactobacillus plantarum MBSa4 isolated from Brazilian salami. Food Control 2016, 60, 103–112. [Google Scholar] [CrossRef] [Green Version]
  37. Barbosa, M.S.; Todorov, S.D.; Belguesmia, Y.; Choiset, Y.; Rabesona, H.; Ivanova, I.V.; Chobert, J.M.; Haertle, T.; Franco, B.D.G.M. Purification and characterization of the bacteriocin produced by Lactobacillus sakei MBSa1 isolated from Brazilian salami. J. Appl. Microbiol. 2014, 116, 1195–1208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Barbosa, M.S.; Todorov, S.D.; Jurkiewicz, C.H.; Franco, B.D.G.M. Bacteriocin production by Lactobacillus curvatus MBSa2 entrapped in calcium alginate during ripening of salami for control of Listeria monocytogenes. Food Control 2015, 47, 147–153. [Google Scholar] [CrossRef]
  39. Todorov, S.D.; Koep, K.S.C.; Van Reenen, C.A.; Hoffman, L.C.; Slinde, E.; Dicks, L.M.T. Production of salami from beef, horse, mutton, blesbok (Damaliscus dorcas phillipsi) and springbok (Antidorcas marsupialis) with bacteriocinogenic strains of Lactobacillus plantarum and Lactobacillus curvatus. Meat Sci. 2017, 77, 405–412. [Google Scholar] [CrossRef]
  40. Enan, G.; El-Essawy, A.A.; Uyttendaele, M.; Debevere, J. Antibacterial activity of Lactobacillus plantarum UG1 isolated from dry sausage: Characterization, production and bactericidal action of plantaricin UG1. Int. J. Food Microbiol. 1996, 30, 189–215. [Google Scholar] [CrossRef]
  41. Hugas, M. Bacteriocinogenic lactic acid bacteria for the biopreservation of meat and meat products. Meat Sci. 1998, 49, S139–S150. [Google Scholar] [CrossRef]
  42. Messi, P.; Bondi, M.; Sabia, C.; Battini, R.; Manicardi, G. Detection and preliminary characterization of a bacteriocin (plantaricin 35d) produced by a Lactobacillus plantarum strain. Int. J. Food Microbiol. 2001, 64, 193–198. [Google Scholar] [CrossRef]
  43. Rattanachaikunsopon, P.; Phumkhachorn, P. Isolation and preliminary characterization of a bacteriocin produced by Lactobacillus plantarum N014 isolated from Nham, a traditional Thai fermented pork. J. Food Protect. 2006, 69, 1937–1943. [Google Scholar] [CrossRef] [Green Version]
  44. Todorov, S.D.; Ho, P.; Vaz-Velho, M.; Dicks, L.M.T. Characterization of bacteriocins produced by two strains of Lactobacillus plantarum isolated from beloura and chouriço, traditional pork products from Portugal. Meat Sci. 2010, 84, 334–343. [Google Scholar] [CrossRef]
  45. Todorov, S.D. Diversity of bacteriocinogenic lactic acid bacteria isolated from boza, a cereal-based fermented beverage from Bulgaria. Food Control 2010, 21, 1011–1021. [Google Scholar] [CrossRef]
  46. Todorov, S.D.; Franco, B.D.G.M.; Vaz-Velho, M. Bacteriocin producing lactic acid bacteria from and for production of salami-like products. A review. Int. Rev. Food Sci. Technol. 2009, 1, 57–61. [Google Scholar]
  47. Todorov, S.D.; von Mollendorff, J.W.; Moelich, E.; Muller, N.; Witthuhn, R.C.; Dicks, L.M.T. Evaluation of potential probiotic properties of Enterococcus mundtii, its survival in boza and in situ bacteriocin production. Food Technol. Biotechnol. 2009, 47, 178–191. [Google Scholar]
  48. Todorov, S.D.; Holzapfel, W.H.; Nero, L.A. Safety evaluation and bacteriocinogenic potential of Pediococcus acidilactici strains isolated from artisanal cheeses. LWT–Food Sci. Technol. 2021, 139, 110550. [Google Scholar] [CrossRef]
  49. Vaz-Velho, M.; Jacome, S.; Noronha, L.; Todorov, S.; Fonseca, S.; Pinheiro, R.; Morais, A.; Silva, J.; Teixeira, P. Comparison of antilisterial effects of two strains of lactic acid bacteria during processing and storage of a Portuguese salami-like product “Alheira”. Chem. Engin. Transact. 2013, 32, 1807–1812. [Google Scholar] [CrossRef]
  50. Wachsman, M.B.; Castilla, V.; de Ruiz Holgado, A.P.; de Torres, R.A.; Sesma, F.; Coto, C.E. Enterocin CRL35 inhibits late stages of HSV-1 and HSV-2 replication in vitro. Antiviral. Res. 2003, 58, 17–24. [Google Scholar] [CrossRef]
  51. Todorov, S.D.; Franco, B.D.G.M.; Wiid, I.J. In vitro study of beneficial properties and safety of lactic acid bacteria isolated from Portuguese fermented meat products. Beneficial Microbs. 2013, 5, 351–366. [Google Scholar] [CrossRef]
  52. Cavicchioli, V.Q.; de Carvalho, O.V.; de Paiva, J.C.; Todorov, S.D.; Júnior, A.S.; Nero, L.A. Inhibition of Herpes simplex virus 1 and Poliovirus (PV 1-1) by bacteriocins from Lactococcus lactis subsp. lactis and Enterococcus durans strains isolated from goat milk. Int. J. Antimicrob. Agents. 2018, 51, 33–37. [Google Scholar] [CrossRef]
  53. Carneiro, B.M.; Braga, A.C.S.; Batista, M.N.; Rahal, P.; Favaro, L.; Penna, A.L.B.; Todorov, S.D. Lactobacillus plantarum ST202Ch and Lactobacillus plantarum ST216Ch—What are the limitations for application? J. Nutr. Health Food Engin. 2014, 1, 00010. [Google Scholar]
  54. Horvath, S. Cytotoxicity of drugs and diverse chemical agents to cell cultures. Toxicol. 1980, 16, 59–66. [Google Scholar] [CrossRef]
  55. Niles, A.L.; Moravec, R.A.; Riss, T.L. Update on in vitro cytotoxicity assays for drug development. Expert. Opin. Drug. Discov. 2008, 3, 655–669. [Google Scholar] [CrossRef] [PubMed]
  56. Myers, M. Direct measurement of cell numbers in microtitre plate cultures using the fluorescent dye SYBR green I. J. Immunol. Methods. 1998, 212, 99–103. [Google Scholar] [CrossRef]
  57. Cook, J.; Mitchell, J. Viability measurements in mammalian cell systems. Anal. Biochem. 1989, 179, 1–7. [Google Scholar] [CrossRef]
  58. Borenfruend, E.; Puerner, J. Toxicity determined in vitro by morphological alterations and neutral red absorption. Toxicol Lett. 1985, 24, 119–124. [Google Scholar] [CrossRef]
  59. Gonzalez-Nicolini, V.; Fux, C.; Fussenegger, M. A novel mammalian cell-based approach for the discovery of anticancer drugs with reduced cytotoxicity on non-dividing cells. Invest. New Drugs. 2004, 22, 253–262. [Google Scholar] [CrossRef]
  60. Cox, C.R.; Coburn, P.S.; Gilmore, M.S. Enterococcal cytolysin: A novel two component peptide system that serves as a bacterial defense against eukaryotic and prokaryotic cells. Curr. Protein Pept. Sci. 2005, 6, 77–84. [Google Scholar] [CrossRef]
  61. Papo, N.; Shai, Y. Host defense peptides as new weapons in cancer treatment. Cell. Mol. Life. Sci. 2005, 62, 784–790. [Google Scholar] [CrossRef]
  62. Hoskin, D.W.; Ramamoorthy, A. Studies on anticancer activities of antimicrobial peptides. BBA Biomembr. 2008, 1778, 357–375. [Google Scholar] [CrossRef] [Green Version]
  63. Kaur, S.; Kaur, S. Bacteriocins as potential anticancer agents. Front. Pharmacol. 2015, 6, 272. [Google Scholar] [CrossRef] [Green Version]
  64. Frazer, A.; Sharratt, M.; Hickman, J. The biological effects of food additives. I. Nisin. J. Sci. Food Agr. 1962, 13, 32–42. [Google Scholar] [CrossRef]
  65. Vaucher, R.d.A.; Velho Gewehr, C.d.C.; Correa, A.P.; Sant’Anna, V.; Ferreira, J.; Brandelli, A. Evaluation of the immunogenicity and in vivo toxicity of the antimicrobial peptide P34. Int. J. Pharm. 2011, 421, 94–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sahoo, T.K.; Jena, P.K.; Prajapati, B.; Gehlot, L.; Patel, A.K.; Seshadri, S. In Vivo assessment of immunogenicity and toxicity of the bacteriocin TSU4 in BALB/c mice. Prob. Antimicrob. Prot. 2017, 9, 345–354. [Google Scholar] [CrossRef]
  67. Marlida, Y.; Arnim, A.; Yuherman, Y.; Rusmarilin, H. Toxicity test pediocin N6 powder produced from isolates Pediococcus Pentosaceus strain N6 on white mice. J. Food Pharm. Sci. 2016, 4, 12–16. [Google Scholar] [CrossRef]
  68. Dover, S.E.; Aroutcheva, A.A.; Faro, S.; Chikindas, M.L. Safety study of an antimicrobial peptide lactocin 160, produced by the vaginal Lactobacillus rhamnosus. Infect. Dis. Obstet. Gynecol. 2007, 2007, 78248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Cebrian, R.; Rodrıguez-Cabezas, M.E.; Martın-Escolano, R.; Rubino, S.; Carrido-Barros, M.; Montalban-Lopez, M.; Rosales, M.J.; Sanchez-Moreno, M.; Valdivia, E.; Martinez-Bueno, M.; et al. Preclinical studies of toxicity and safety of the AS-48 bacteriocin. J. Adv. Res. 2019, 20, 129–139. [Google Scholar] [CrossRef]
  70. Murinda, S.; Rashid, K.; Roberts, R. In vitro assessment of the cytotoxicity of nisin, pediocin, and selected colicins on simian Virus 40-transfected human colon and Vero monkey kidney cells with trypan blue staining viability assays. J. Food Prot. 2003, 66, 847–853. [Google Scholar] [CrossRef] [PubMed]
  71. Paiva, A.D.; de Oliveira, M.D.; de Paula, S.O.; Baracat-Pereira, M.C.; Breukink, E.; Mantovani, H.C. Toxicity of bovicin HC5 against mammalian cell lines and the role of cholesterol in bacteriocin activity. Microbiolology 2012, 158, 2851–2858. [Google Scholar] [CrossRef]
  72. Favaro, L.; Todorov, S.D. Bacteriocinogenic LAB strains for fermented meat preservation: Perspectives, challenges and limitations. Prob. Antimicrob. Prot. 2017, 9, 444–458. [Google Scholar] [CrossRef]
  73. Cleveland, J.; Montville, T.J.; Nes, I.F.; Chikindas, M.L. Bacteriocins: Safe, natural antimicrobials for food preservation. Int. J. Food Microbiol. 2001, 71, 1–20. [Google Scholar] [CrossRef]
  74. De Vuyst, L.; Leroy, F. Bacteriocins from lactic acid bacteria: Production, purification, and food applications. J. Mol. Microb. Biotech. 2007, 13, 194–199. [Google Scholar] [CrossRef]
  75. Fernandez, B.; Le Lay, C.; Jean, J.; Fliss, I. Growth, acid production and bacteriocin production by probiotic candidates under simulated colonic conditions. J. Appl. Microbiol. 2013, 114, 877–885. [Google Scholar] [CrossRef] [PubMed]
  76. Kheadr, E.; Zihler, A.; Dabour, N.; Lacroix, C.; Le Blay, G.; Fliss, I. Study of the physicochemical and biological stability of pediocin PA-1 in the upper gastrointestinal tract conditions using a dynamic in vitro model. J. Appl. Microbiol. 2010, 109, 54–64. [Google Scholar] [CrossRef]
  77. Birri, D.J.; Brede, D.A.; Nes, I.F. Salivaricin D, a novel intrinsically trypsin-resistant lantibiotic from Streptococcus salivarius 5M6c isolated from a healthy infant. Appl. Environ. Microbiol. 2012, 78, 402–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Johnson, E.M.; Jung, D.Y.; Jin, D.Y.; Jayabalan, D.R.; Yang, D.S.H.; Suh, J.W. Bacteriocins as food preservatives: Challenges and emerging horizons. Crit. Rev. Food Sci. Nutr. 2018, 58, 2743–2767. [Google Scholar] [CrossRef]
  79. O’Shea, E.F.; O’Connor, P.M.; Cotter, P.D.; Ross, P.P.; Hill, C. Synthesis of trypsinresistant variants of the Listeria-active bacteriocin salivaricin P. Appl. Environ. Microbiol. 2010, 76, 5356–5362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Rollema, H.S.; Kuipers, O.P.; Both, P.; de Vos, W.M.; Siezen, R.J. Improvement of solubility and stability of the antimicrobial peptide nisin by protein engineering. Appl. Environ. Microbiol. 1995, 61, 2873–2878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Field, D.; Blake, T.; Mathur, H.; O’Connor, P.M.; Cotter, P.D.; Ross, P.R.; Hill, C. Bioengineering nisin to overcome the nisin resistance protein. Mol. Microbiol. 2019, 111, 717–731. [Google Scholar] [CrossRef]
  82. Gomaa, A.I.; Martinent, C.; Hammami, R.; Fliss, I.; Subirade, M. Dual coating of liposomes as encapsulating matrix of antimicrobial peptides: Development and characterization. Front. Chem. 2017, 5, 103. [Google Scholar] [CrossRef] [Green Version]
  83. Habib, W.; Sakr, A. Development and human in vivo evaluation of a colonic drug delivery system. Pharm. Ind. 1999, 61, 1145–1149. [Google Scholar]
  84. Gough, R.; Cabrera Rubio, R.; O’Connor, P.M.; Crispie, F.; Brodkorb, A.; Miao, S.; Hill, C.; Ross, R.P.; Cotter, P.D.; Nilaweera, K.N.; et al. Oral delivery of nisin in resistant starch-based matrices alters the gut microbiota in mice. Front. Microbiol. 2018, 9, 1186. [Google Scholar] [CrossRef] [Green Version]
  85. De Groot, A.S.; Scott, D.W. Immunogenicity of protein therapeutics. Trends Immunol. 2007, 28, 482–490. [Google Scholar] [CrossRef] [PubMed]
  86. Lohans, C.T.; Vederas, J.C. Development of class IIa bacteriocins as therapeutic agents. Int. J. Microbiol. 2011, 2012, 386410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Bhunia, A.K.; Johnson, M.C.; Ray, B.; Belden, E.L. Antigenic properties of pediocin AcH produced by Pediococcus acidilactici H. J. Appl. Microbiol. 1990, 69, 211–215. [Google Scholar] [CrossRef]
  88. Pablo, M.A.; Gaforio, J.J.; Gallego, A.M.; Ortega, E.; Galvez, A.M.; de Cienfuegos Lopez, G.A. Evaluation of immunomodulatory effects of nisin-containing diets on mice. FEMS Immunol. Med. Micro. 1999, 24, 35–42. [Google Scholar] [CrossRef] [Green Version]
  89. Scholl, D.; Martin, D.W. Antibacterial efficacy of R-type pyocins towards Pseudomonas aeruginosa in a murine peritonitis model. Antimicrob. Agents Chemother. 2008, 52, 1647–1652. [Google Scholar] [CrossRef] [Green Version]
  90. McCaughey, L.C.; Ritchie, N.D.; Douce, G.R.; Evans, T.J.; Walker, D. Efficacy of species-specific protein antibiotics in a murine model of acute Pseudomonas aeruginosa lung infection. Sci. Rep. 2016, 6, 30201. [Google Scholar] [CrossRef] [Green Version]
  91. Belguesmia, Y.; Naghmouchi, K.; Chihib, N.E.; Drider, D. Class IIa bacteriocins: Current knowledge and perspectives. In Prokaryotic Antimicrobial Peptides; Drider, D., Rebuffat, S., Eds.; Springer: New York, NY, USA, 2011. [Google Scholar] [CrossRef]
  92. Fujiwara, S.; Hashiba, H.; Hirota, T.; Forstner, J.F. Inhibition of the binding of enterotoxigenic Escherichia coli Pb176 to human intestinal epithelial cell line HCT-8 by an extracellular protein fraction containing BIF of Bifidobacterium longum SBT2928: Suggestive evidence of blocking of the binding receptor gangliotetraosylceramide on the cell surface. Int. J. Food Microbiol. 2001, 67, 97–106. [Google Scholar] [CrossRef]
  93. Fujiwara, S.; Hashiba, H.; Hirota, T.; Forstner, J.F. Proteinaceous factor(s) in culture supernatant fluids of bifidobacteria which prevents the binding of enterotoxigenic Escherichia coli to gangliotetraosylceramide. Appl. Environ. Microbiol. 1997, 63, 506–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Di Cagno, R.; De Angelis, M.; Calasso, M.; Vincentini, O.; Vernocchi, P.; Ndagijimana, M.; De Vincenzi, M.; Dessì, M.R.; Guerzoni, M.E.; Gobbetti, M. Quorum sensing in sourdough Lactobacillus plantarum DC400: Induction of plantaricin A (PlnA) under co-cultivation with other lactic acid bacteria and effect of PlnA on bacterial and Caco-2 cells. Proteomics 2010, 10, 2175–2190. [Google Scholar] [CrossRef] [Green Version]
  95. Villarante, K.I.; Elegado, F.B.; Iwatani, S.; Zendo, T.; Sonomoto, K.; de Guzman, E.E. Purification, characterization and in vitro cytotoxicity of the bacteriocin from Pediococcus acidilactici K2a2–3 against human colon adenocarcinoma (HT29) and human cervical carcinoma (HeLa) cells. World J. Microbiol. Biotechnol. 2011, 27, 975–980. [Google Scholar] [CrossRef]
  96. Olejnik-Schmidt, A.K.; Schmidt, M.T.; Sip, A.; Szablewski, T.; Grajek, W. Expression of bacteriocin divercin AS7 in Escherichia coli and its functional analysis. Ann. Microbiol. 2014, 64, 1197–1202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Aasen, I.M.; Møretrø, T.; Katla, T.; Axelsson, L.; Storrø, I. Influence of complex nutrients, temperature and pH on bacteriocin production by Lactobacillus sakei CCUG 42687. Appl. Microbiol. Biotechnol. 2000, 53, 159–166. [Google Scholar] [CrossRef] [PubMed]
  98. Bogovic-Matijasic, B.; Rogelj, I. Bacteriocin complex of Lactobacillus acidophilus LF221—Production studies in MRS-media at different pH-values and effect against Lactobacillus helveticus ATCC 15009. Process Biochem. 1988, 33, 345–352. [Google Scholar] [CrossRef]
  99. Krier, F.; Revol-Junelles, A.M.; Germain, P. Influence of temperature and pH on production of two bacteriocins by Leuconostoc mesenteroides subsp. mesenteroides FR52 during batch fermentation. Appl. Microbiol. Biotechnol. 1998, 50, 359–363. [Google Scholar] [CrossRef] [PubMed]
  100. Matsusaki, H.; Endo, N.; Sonomoto, K.; Ishizaki, A. Lantibiotic nisin Z fermentative production by Lactococcus lactis IO-1: Relationship between production of the lantibiotic and lactate and cell growth. Appl. Microbiol. Biotechnol. 1996, 45, 36–40. [Google Scholar] [CrossRef]
  101. Mortvedt-Abildgaa, C.I.; Nissen-Meyer, J.; Jelle, B.; Grenov, B.; Skaugen, M.; Nes, I.F. Production and pH-dependent bactericidal activity of lactocin S, a lantibiotic from Lactobacillus sake L45. Appl. Environ. Microbiol. 1995, 61, 175–179. [Google Scholar] [CrossRef] [Green Version]
  102. Drider, D.; Fimland, G.; Héchard, Y.; McMullen, L.M.; Prévost, H. The continuing story of class IIa bacteriocins. Microbiol. Mol. Biol. Rev. 2006, 70, 564–582. [Google Scholar] [CrossRef] [Green Version]
  103. Soliman, W.; Bhattacharjee, S.; Kaur, K. Adsorption of an antimicrobial peptide on self-assembled monolayers by molecular dynamics simulation. J. Phys. Chem. B. 2010, 114, 11292–11302. [Google Scholar] [CrossRef]
  104. Grein, F.; Schneider, T.; Sahl, H.G. Docking on lipid II-A widespread mechanism for potent bactericidal activities of antibiotic peptides. J. Mol. Biol. 2019, 431, 3520–3530. [Google Scholar] [CrossRef]
  105. Wang, X.; Gu, Q.; Breukink, E. Non-lipid II targeting lantibiotics. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183244. [Google Scholar] [CrossRef]
  106. Zhu, L.; Zeng, J.; Wang, C.; Wang, J. Structural basis of pore formation in the mannose phosphotransferase system by pediocin PA-1. Appl. Environ. Microbiol. 2022, 88, e0199221. [Google Scholar] [CrossRef] [PubMed]
  107. Gadea, R.; Glibota, N.; Perez Pulido, R.; Galvez, A.; Ortega, E. Effects of exposure to biocides on susceptibility to essential oils and chemicals preservatives in bacteria from organic foods. Food Control 2017, 80, 176–182. [Google Scholar] [CrossRef]
  108. Fernandez-Fuentes, M.A.; Ortega Morente, E.; Abriouel, H.; Perez Pulido, R.; Galvez, A. Isolation and identification of bactéria from organic foods: Sensitivity to biocides and antibiotics. Food Control 2012, 26, 73–78. [Google Scholar] [CrossRef]
  109. Ribeiro, A.M.; Paiva, A.D.; Cruz, A.M.; Vanetti, M.C.; Ferreira, S.O.; Mantovani, H.C. Bovicin HC5 and nisin reduce cell viability and the thermal resistance of Alicyclobacillus acidoterrestris endospores in fruit juices. J. Sci. Food Agric. 2022, 102, 3994–4002. [Google Scholar] [CrossRef] [PubMed]
  110. De Carvalho, A.A.T.; Montovani, H.C.; Vanetto, M.C.D. Bactericidal effect of bovicin HC5 and nisin against Clostridium tyrobutyricum isolated from spoiled mango pulp. Lett. Appl. Microbiol. 2007, 45, 68–74. [Google Scholar] [CrossRef] [Green Version]
  111. Aymerich, T.; Jofré, A.; Bover-Cid, S. Enterocin A-based antimicrobial film exerted strong antilisterial activity in sliced dry-cured ham immediately and after 6 months at 8 °C. Food Microbiol. 2022, 105, 104005. [Google Scholar] [CrossRef] [PubMed]
  112. Xiang, Y.-Z.; Wu, G.; Zhang, Y.-P.; Yang, L.-Y.; Zhang, Y.-M.; Zhao, Z.-S.; Deng, X.-Y.; Zhang, Q.-L. Inhibitory effect of a new bacteriocin RSQ04 purified from Lactococcus lactis on Listeria monocytogenes and its application on model food system. LWT Food Sci. Technol. 2022, 164, 113626. [Google Scholar] [CrossRef]
  113. Doshi, M.N.; Nair, K.; Hussan, U.; Jaoua, S. Pyocin QDD1: A highly thermostable bacteriocin produced by Pseudomonas aeruginosa QDD1 for the biocontrol of foodborne pathogens Staphylococcus aureus and Bacillus cereus. Biores. Technol. Rep. 2022, 18, 101106. [Google Scholar] [CrossRef]
  114. Zhang, Y.-M.; Jiang, Y.-H.; Li, H.-W.; Zhang, Q.-L. Purification and characterization of Lactobacillus plantarum—Derivated bacteriocin with activity against Staphylococcus argenteus planktonic cells and biofilm. J. Food Sci. 2022, 87, 2718–2731. [Google Scholar] [CrossRef]
  115. Valledor, S.J.D.; Dioso, C.M.; Bucheli, J.E.V.; Park, Y.J.; Suh, D.H.; Jung, E.S.; Kim, B.; Holzapfel, W.H.; Todorov, S.D. Characterization and safety evaluation of two beneficial, enterocin-producing Enterococcus faecium strains isolated from kimchi, a Korean fermented cabbage. Food Microbiol. 2022, 102, 103886. [Google Scholar] [CrossRef]
  116. Jiang, Y.H.; Xin, W.G.; Yang, L.Y.; Ying, J.P.; Zhao, Z.S.; Lin, L.B.; Li, X.Z.; Zhang, Q.L. A novel bacteriocin against Staphylococcus aureus from Lactobacillus paracasei isolated from Yunnan traditional fermented yogurt: Purification, antibacterial characterization, and antibiofilm activity. J. Dairy Sci. 2022, 105, 2094–2107. [Google Scholar] [CrossRef] [PubMed]
  117. Serra-Castello, C.; Costa, J.C.C.P.; Jofre, A.; Bolivar, A.; Perez-Rodriguez, F.; Bover-Cid, S. A mathematical model to predict the antilisterail bioprotective effect of Latilactobacillus sakei CTC494 in vacuum packed cooked ham. Int. J. Food Microbiol. 2022, 363, 109491. [Google Scholar] [CrossRef]
  118. Zadeh, R.G.; Asgharzadeh, S.; Darbandi, A.; Aliramezani, A.; Jazi, F.M. Characterization of bacteriocins produced by Lactobacillus species against adhesion and invasion of Listeria monocytogenes isolated from different samples. Microbial. Pathogen. 2022, 162, 105307. [Google Scholar] [CrossRef] [PubMed]
  119. Fugaban, J.I.I.; Holzapfel, W.H.; Todorov, S.D. Probiotic potential and safety assessment of bacteriocinogenic Enterococcus faecium strains with antibacterial activity against Listeria and vancomycin-resistant enterococci. Curr. Res. Microb. Sci. 2021, 30, 100070. [Google Scholar] [CrossRef]
  120. Abitayeva, G.K.; Urazova, M.S.; Abilkhadirov, A.S.; Sarmurzina, Z.S.; Shaikhin, S.M. Characterization of a new bacteriocin-like inhibitory peptide produced by Lactobacillus sakei B-RKM 0559. Biotechnol. Lett. 2021, 43, 2243–2257. [Google Scholar] [CrossRef] [PubMed]
  121. Han, J.; Meng, X.; Shen, H.; Luo, W.; Yao, S.; Yang, J.; Zhu, Q.; Tian, Y.; Wang, S. Purification, molecular characterization of Lactocin 63 produced by Lactobacillus coryniformis FZU63 and its antimicrobial mode of action against Shewanella putrefaciens. Appl. Microbiol. Biotechnol. 2021, 105, 6921–6930. [Google Scholar] [CrossRef]
  122. Barbosa, J.; Albano, H.; Silva, B.; Almeida, M.H.; Nogueira, T.; Teixeira, P. Characterization of a Lactiplantibacillus plantarum R23 Isolated from Arugula by whole-genome sequencing and its bacteriocin production ability. Int. J. Environ. Res. Public Health. 2021, 18, 5515. [Google Scholar] [CrossRef]
  123. Sheoran, P.; Tiwari, S.K. Synergistically-acting enterocin LD3 and plantaricin LD4 against Gram-positive and Gram-negative pathogenic bacteria. Probiot. Antimicro. Prot. 2021, 13, 542–554. [Google Scholar] [CrossRef]
  124. Yan, H.; Lu, Y.; Li, X.; Yi, Y.; Wang, X.; Shan, Y.; Liu, B.; Zhou, Y.; Lu, X. Action mode of bacteriocin BM1829 against Escherichia coli and Staphylococcus aureus. Food Biosci. 2021, 39, 100794. [Google Scholar] [CrossRef]
  125. Alang, H.; Kusnadi, J.; Ardyati, T.; Suharjono. Optimisation and characterization of enterocin Enterococcus faecalis K2B1 isolated from Toraja’s Belang buffalo milk, South Sulawesi, Indonesia. Biodiversitas J. Biol. Divers. 2020, 21, 1236–1242. [Google Scholar] [CrossRef]
  126. Ananou, S.; Lotfi, S.; Azdad, O.; Nzoyikorera, N. Production, recovery and characterization of an enterocin with anti-listerial activity produced by Enterococcus hirae OS1. Appl. Food Biotechnol. 2020, 7, 103–114. [Google Scholar] [CrossRef]
  127. Fugaban, J.I.I.; Bucheli, J.E.V.; Kim, B.; Holzapfel, W.H.; Todorov, S.D. Safety and beneficial properties of bacteriocinogenic Pediococcus acidilactici and Pediococcus pentosaceus isolated from silage. Lett. Appl. Microbiol. 2021, 73, 725–734. [Google Scholar] [CrossRef]
  128. De Souza, B.M.S.; Borgonovi, T.F.; Casaroti, S.N.; Todorov, S.D.; Penna, A.L.B. Lactobacillus casei and Lactobacillus fermentum strains isolated from Mozzarella cheese: Probiotic potential, safety, acidification kinetic parameters and viability under gastrointestinal tract conditions. Prob. Antimicrob. Prot. 2019, 11, 382–396. [Google Scholar] [CrossRef] [PubMed]
  129. Casarotti, S.N.; Penna, A.L.B. Acidification profile, probiotic in vitro gastrointestinal tolerance and viability in fermented milk with fruit flours. Int. Dairy J. 2015, 41, 1–6. [Google Scholar] [CrossRef]
  130. Worsztynowicz, P.; Bialas, W.; Grajek, W. Integrated approach for obtaining bioactive peptides from whey proteins hydrolysedusing a new proteolytic lactic acid bacteria. Food Chem. 2020, 312, 126035. [Google Scholar] [CrossRef]
  131. Garcia-Cano, I.; Rocha-Mendoza, D.; Ortega-Anaya, J.; Wang, K.; Kosmerl, E.; Jimenez-Flores, R. Lactic acid bactéria isolated from dairy products as potential producers of lipolytic, proteolytic and antibacterial proteins. Appl. Microbiol. Biotechnol. 2019, 103, 5243–5257. [Google Scholar] [CrossRef] [Green Version]
  132. Biscola, V.; Choiset, Y.; Rabesona, H.; Chobert, J.-M.; Haertle, T.; Franco, B.D.G.M. Brazilian artisanal ripened cheeses as source of proteolityc lactic acid bacteria capable of reducing cow milk allergy. J. Appl. Microbiol. 2018, 125, 564–574. [Google Scholar] [CrossRef]
  133. Brown, L.; Pingitore, E.V.; Mozzi, F.; Saavedra, L.; Villegas, J.M.; Hebert, E.M. Lactic acid bacteria as cell factories for the generation of bioactive peptides. Prot. Pept. Lett. 2017, 24, 146–155. [Google Scholar] [CrossRef]
  134. Herran, A.R.; Perez-Andres, J.; Caminero, A.; Nistal, E.; Vivas, S.; Ruiz de Morales, J.M.; Casqueiro, J. Gluten-degrading bacteria are present in the human small intestine of healthy volunteers and celic patiens. Res. Microbiol. 2017, 168, 673–684. [Google Scholar] [CrossRef]
  135. Perin, L.M.; Belviso, S.; Dal Bello, B.; Nero, L.A.; Cocolin, L. Technological properties and biogenic amines production by bacteriocinogenic lactococci and enterococci strains isolated from raw goat’s milk. J. Food Protect. 2017, 80, 151–157. [Google Scholar] [CrossRef]
  136. Perin, L.M.; Miranda, R.P.; Todorov, S.D.; Franco, B.D.G.M.; Nero, L.A. Virulence, antibiotic resistenceand biogenic amines of bacteriocinogenic lactococci and enterococci isolated from goat milk. Int. J. Food Microbiol. 2014, 185, 121–126. [Google Scholar] [CrossRef] [PubMed]
  137. Suvorov, A. What is wrong with enterococcal probiotics? Prob. Antimicrob. Prot. 2020, 12, 1–4. [Google Scholar] [CrossRef] [PubMed]
  138. Todorov, S.D.; Holzapfel, W.H. Traditional cereal fermented foods as sources of functional microorganisms. In Advances in Fermented Foods and Beverages: Improving Quality, Technologies and Health Benefits; Holzapfel, W.H., Ed.; Woodhead Publishing: London, UK, 2013; pp. 123–153. [Google Scholar]
  139. Von Mollendorff, J.; Todorov, S.D.; Vaz-Velho, M. Cereal-based fermented foods are rich source of probiotics and bacteriocin producer lactic acid bacteria. In Functional Properties of Traditional Foods; ISEKI Food Series Volume 12; Kristbergsson, K., Ötles, S., Eds.; Springer Publishing Group: New York, NY, USA, 2016; pp. 157–188. [Google Scholar] [CrossRef]
  140. Öncül, N.; Yıldırım, Z. Inhibitory effect of bacteriocins against Escherichia coli O157:H7. Food Sci. Technol. Int. 2019, 25, 504–514. [Google Scholar] [CrossRef] [PubMed]
  141. Gök, C.M.; Özden, T.B.; Akpinar, K.D.; Tuncer, Y. Bacteriocinogenic properties and safety evaluation of Enterococcus faecium YT52 isolated from boza, a traditional cereal based fermented beverage. J. Verbrauch. Lebensm. 2019, 14, 41–53. [Google Scholar] [CrossRef]
  142. Koral, G.; Tuncer, Y. Nisin Z—Producing Lactococcus lactis subsp. lactis GYI32 isolated from boza. J. Food Proc. Preserv. 2014, 38, 1044–1053. [Google Scholar] [CrossRef]
  143. Botes, A.; Todorov, S.D.; von Mollendorff, J.W.; Botha, A.; Dicks, L.M.T. Identification of Lactic Acid Bacteria and Yeast from Boza. Process Biochem. 2007, 42, 267–270. [Google Scholar] [CrossRef]
  144. Todorov, S.D.; Botes, M.; Guigas, C.; Schillinger, U.; Wiid, I.; Wachsman, M.B.; Holzapfel, W.H.; Dicks, L.M.T. Boza, a natural source of probiotic lactic acid bacteria. J. Appl. Microbiol. 2008, 104, 465–477. [Google Scholar] [CrossRef]
  145. Von Mollendorff, J.W.; Todorov, S.D.; Dicks, L.M.T. Factors affecting the adsorption of bacteriocins to Latobacillus sakei and Enterococcus sp. Appl. Biochem. Biotechnol. 2007, 142, 209–220. [Google Scholar] [CrossRef]
  146. Pingitore, E.V.; Todorov, S.D.; Sesma, F.; Franco, B.D.G.M. Application of bacteriocinogenic Enterococcus mundtii CRL35 and Enterococcus faecium ST88Ch in the control of Listeria monocytogenes in fresh Minas cheese. Food Microbiol. 2012, 32, 38–47. [Google Scholar] [CrossRef]
  147. Martinez, R.C.R.; Staliano, C.D.; Vieira, A.D.S.; Villarreal, M.L.M.; Todorov, S.D.; Saad, S.M.I.; Franco, B.D.G.M. Bacteriocin production and inhibition of Listeria monocytogenes by Lactobacillus sakei subsp. sakei 2a in a potentially synbiotic cheese spread. Food Microbiol. 2015, 48, 143–152. [Google Scholar] [CrossRef]
  148. Schirru, S.; Favaro, L.; Mangia, N.P.; Basaglia, M.; Casella, S.; Comunian, R.; Fancello, F.; Franco, B.D.G.M.; Oliveira, R.P.S.; Todorov, S.D. Comparison of bacteriocins production from Enterococcus faecium strains in cheese whey and optimised commercial MRS medium. Ann. Microbiol. 2014, 64, 321–331. [Google Scholar] [CrossRef]
  149. Heng, N.C.K.; Wescombe, P.A.; Burton, J.P.; Jack, R.W.; Tagg, J.R. The diversity of bacteriocins in Gram-positive bacteria. In Bacteriocins; Riley, M.A., Chavan, M.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar] [CrossRef]
  150. Jozala, A.F.; de Andrade, M.S.; de Arauz, L.J.; Pessoa, A.; Penna, T.C.V. Nisin production utilizing skimmed milk aiming to reduce process cost. Appl. Biochem. Biotechnol. 2007, 137, 515. [Google Scholar] [CrossRef] [PubMed]
  151. Pongtharangkul, T.; Demirci, A. Evaluation of agar diffusion bioassay for nisin quantification. Appl. Microbiol. Biotechnol. 2004, 65, 268–272. [Google Scholar] [CrossRef] [PubMed]
  152. Reis, J.A.; Paula, A.T.; Casarotti, S.N.; Penna, A.L.B. Lactic acid bacteria antimicrobial compounds: Characteristics and applications. Food Eng. Rev. 2012, 4, 124–140. [Google Scholar] [CrossRef]
  153. Corredoira, J.; Coira, A.; Alonso, M.A.; Varela, J. Association between Streptococcus infantarius (Formerly, S. bovis II/1). J. Clin. Microbiol. 2008, 46, 1570. [Google Scholar] [CrossRef] [Green Version]
  154. Boleij, A.; Gelder, M.H.J.; Swinnkels, D.W.; Tjalsma, H. Clinical importance of Streptococcus gallolyticus infection among colorectal cancer patients: Systematic review and meta-analysis. Clin. Infect. Dis. 2011, 53, 870–878. [Google Scholar] [CrossRef] [Green Version]
  155. Jans, C.; Kaindi, D.W.M.; Böck, D.; Njage, P.M.K.; Kouamé-Sina, S.M.; Bonfoh, B.; Lacroix, C.; Meile, L. Prevalence and comparison of Streptococcus infantarius subsp. infantarius and Streptococcus gallolyticus subsp. macedonicus in raw and fermented dairy products from East and West Africa. Int. J. Food Microbiol. 2013, 167, 186–195. [Google Scholar] [CrossRef] [Green Version]
  156. Todorov, S.D.; Dioso, C.M.; Liong, M.-T.; Vasileva, T.; Moncheva, P.; Ivanova, I.V.; Iliev, I. Lukanka, a semi-dried fermented traditional Bulgarian sausage: Role of the bacterial cultures in its technological, safety and beneficial characteristics. Appl. Food Biotechnol. 2022, 9, 255–265. [Google Scholar] [CrossRef]
  157. Schillinger, U.; Geisen, R.; Holzapfel, W.H. Potential of antagosnistic microorganisms and bacteriocins for the biological preservation of food. Trends Food Sci. Technol. 1996, 7, 158–164. [Google Scholar] [CrossRef]
  158. Foulquié-Moreno, M.R.; Sarantinopoulos, P.; Tsakalidou, E.; De Vuyst, L. The role and application of enterococci in food and health. Int. J. Food Microbiol. 2006, 106, 1–24. [Google Scholar] [CrossRef]
  159. Landman, D.; Quale, J.M. Management of infections due to resistant enterococci: A review of therapeutic options. J. Antimicrob. Chemother. 1997, 40, 161–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Cetinkaya, Y.; Falk, P.; Mayhall, C.G. Vancomycin-resistant enterococci. Clin. Microbiol. Rev. 2000, 13, 686–707. [Google Scholar] [CrossRef] [PubMed]
  161. Bonten, M.J.; Willems, R.; Weinstein, R.A. Vancomycin-resistant enterococci: Why are they here, and where do they come from? Lancet Infect. Dis. 2001, 1, 314–325. [Google Scholar] [CrossRef]
  162. Franz, C.M.; Stiles, M.E.; Schleifer, K.H.; Holzapfel, W.H. Enterococci in foods—A conundrum for food safety. Int. J. Food Microbiol. 2003, 88, 105–122. [Google Scholar] [CrossRef]
  163. Klare, I.; Konstabel, C.; Badstübner, D.; Werner, G.; Witte, W. Occurrence and spread of antibiotic resistances in Enterococcus faecium. Int. J. Food Microbiol. 2003, 88, 269–290. [Google Scholar] [CrossRef]
  164. Johnson, A.G.; Nguyen, T.V.; Day, R.O. Do nonsteroidal anti-inflammatory drugs affect blood pressure? A meta-analysis. Ann. Intern. Med. 1994, 121, 289–300. [Google Scholar] [CrossRef]
  165. Dos Santos, K.M.O.; de Matos, C.R.; Salles, H.O.; Franco, B.D.G.M.; Arellano, K.; Holzapfel, W.H.; Todorov, S.D. Exploring beneficial/virulence properties of two dairy related strains of Streptococcus infantarius subsp. infantarius. Probiot. Antimicrob. Prot. 2020, 12, 1524–1541. [Google Scholar] [CrossRef]
  166. Nascimento, L.C.S.; Casarotti, S.N.; Todorov, S.D.; Penna, A.L.B. Probiotic potential and safety of enterococci strains. Ann. Microbiol. 2018, 69, 241–252. [Google Scholar] [CrossRef]
  167. Hammes, W.P.; Tichaczek, P.S. The potential of lactic acid bacteria for the production of safe and wholesome food. Z. Lebensm. Unters. Forsch. 1994, 198, 193–201. [Google Scholar] [CrossRef]
  168. Land, M.H.; Rouster-Stevens, K.; Woods, C.R.; Cannon, M.L.; Cnota, J.; Shetty, A.K. Lactobacillus sepsis associated with probiotic therapy. Pediatrics 2005, 115, 178–181. [Google Scholar] [CrossRef]
  169. Kulkarni, H.S.; Khoury, C.C. Sepsis associated with Lactobacillus bacteremia in a patient with ischemic colitis. Indian J. Crit. Care Med. 2014, 18, 606–608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Salminen, M.K.; Rautelin, H.; Tynkkynen, S.; Poussa, T.; Saxelin, M.; Valtonen, V.; Järvinen, A. Lactobacillus bacteremia, clinical significance, and patient outcome, with special focus on probiotic L. rhamnosus GG. Clin. Infect. Dis. 2004, 38, 62–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Karime, C.; Barrios, M.S.; Wiest, N.E.; Stancampiano, F. Lactobacillus rhamnosus sepsis, endocarditis and septic emboli in a patient with ulcerative colitis taking probiotics. BMJ Case Rep. 2022, 15, e249020. [Google Scholar] [CrossRef] [PubMed]
  172. Aguirre, M.; Collins, M.D. Lactic acid bacteria and human clinical infection. J. Appl. Bacteriol. 1993, 75, 95–107. [Google Scholar] [CrossRef] [PubMed]
  173. Kalschne, D.L.; Womer, R.; Mattana, A.; Sarmento, C.M.; Colla, L.M.; Colla, E. Characterization of the spoilage lactic acid bacteria in “sliced vacuum-packed cooked ham”. Braz. J. Microbiol. 2015, 46, 173–181. [Google Scholar] [CrossRef] [Green Version]
  174. Andreevskaya, M.; Jääskeläinen, E.; Johansson, P.; Ylinen, A.; Paulin, L.; Björkroth, J.; Auvinen, P. Food spoilage-associated Leuconostoc, Lactococcus, and Lactobacillus species display different survival strategies in response to competition. Appl. Environ. Microbiol. 2018, 84, e00554-18. [Google Scholar] [CrossRef] [Green Version]
  175. Umu, Ö.C.; Bäuerl, C.; Oostindjer, M.; Pope, P.B.; Hernández, P.E.; Pérez-Martínez, G.; Diep, D.B. The potential of class II bacteriocins to modify gut microbiota to improve host health. PLoS ONE 2016, 11, e0164036. [Google Scholar] [CrossRef] [Green Version]
  176. De Koning, H.P. Drug resistance in protozoan parasites. Emerg. Top. Life Sci. 2017, 1, 627–632. [Google Scholar] [CrossRef]
  177. Uebanso, T.; Shimohata, T.; Mawatari, K.; Takahashi, A. Functional roles of B-vitamins in the gut and gut microbiome. Mol. Nutr. Food Res. 2020, 64, e2000426. [Google Scholar] [CrossRef]
  178. Ammor, M.S.; Flórez, A.B.; Mayo, B. Antibiotic resistance in non-enterococcal lactic acid bacteria and bifidobacteria. Food Microbiol. 2007, 24, 559–570. [Google Scholar] [CrossRef]
  179. Teuber, M.; Meile, L.; Schwarz, F. Acquired antibiotic resistance in lactic acid bacteria from food. In Lactic Acid Bacteria: Genetics, Metabolism and Applications; Konings, W.N., Kuipers, O.P., In’t Veld, J.H.J.H., Eds.; Springer: Dordrecht, Germany, 1919. [Google Scholar] [CrossRef]
  180. Levy, S.B.; Marshall, B. Antibacterial resistance worldwide: Causes, challenges and responses. Nat. Med. 2004, 10 (Suppl. 12), S122–S129. [Google Scholar] [CrossRef] [PubMed]
  181. Bucheli, J.E.V.; Todorov, S.D.; Holzapfel, W.H. Role of gastrointestinal microbial populations, a terra incognita of the human body in the management of intestinal bowel disease and metabolic disorders. Benef. Microbes. 2022, 22, 1–24. [Google Scholar] [CrossRef] [PubMed]
  182. Le Blay, G.; Hammami, R.; Lacroix, C.; Fliss, I. Stability and inhibitory activity of pediocin PA-1 against Listeria sp. in simulated physiological conditions of the human terminal ileum. Prob. Antimicrob. Prot. 2012, 4, 250–258. [Google Scholar] [CrossRef] [PubMed]
  183. Guinane, C.M.; Cotter, P.D.; Hill, C.; Ross, R.P. Spontaneous resistance in Lactococcus lactis IL1403 to the lantibiotic lacticin 3147. FEMS Microbiol. Lett. 2006, 260, 77–83. [Google Scholar] [CrossRef] [Green Version]
  184. Le Lay, C.; Fernandez, B.; Hammami, R.; Ouellette, M.; Fliss, I. On Lactococcus lactis UL719 competitivity and nisin (Nisaplin®) capacity to inhibit Clostridium difficile in a model of human colon. Front. Microbiol. 2015, 6, 1020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Bernbom, N.; Licht, T.R.; Brogren, C.-H.; Jelle, B.; Johansen, A.H.; Badiola, I.; Vogensen, F.K.; Nørrung, B. Effects of Lactococcus lactis on composition of intestinal microbiota: Role of nisin. Appl. Environ. Microbiol. 2006, 72, 239–244. [Google Scholar] [CrossRef] [Green Version]
  186. Dabour, N.; Zihler, A.; Kheadr, E.; Lacroix, C.; Fliss, I. In Vivo study on the effectiveness of pediocin PA-1 and Pediococcus acidilactici UL5 at inhibiting Listeria monocytogenes. Int. J. Food Microbiol. 2009, 133, 225–233. [Google Scholar] [CrossRef]
  187. Rea, M.C.; Dobson, A.; O’Sullivan, O.; Crispie, F.; Fouhy, F.; Cotter, P.D.; Shanahan, F.; Kiely, B.; Hill, C.; Ross, R.P. Effect of broad- and narrow-spectrum antimicrobials on Clostridium difficile and microbial diversity in a model of the distal colon. Proc. Natl. Acad. Sci. USA 2011, 108 (Suppl. 1), 4639–4644. [Google Scholar] [CrossRef] [Green Version]
  188. Blay, G.L.; Lacroix, C.; Zihler, A.; Flies, I. In Vitro inhibition activity of nisin A, nisin Z, pediocin PA-1 and antibiotics against common intestinal bacteria. Lett. Appl. Microbiol. 2007, 45, 252–257. [Google Scholar] [CrossRef]
  189. Dobson, A.; Crispie, F.; Rea, M.C.; O’Sullivan, O.; Casey, P.G.; Lawlor, P.G.; Cotter, P.D.; Ross, P.; Gardiner, G.E.; Hill, C. Fate and efficacy of lacticin 3147-producing Lactococcus lactis in the mammalian gastrointestinal tract. FEMS Microbiol. Ecol. 2011, 76, 602–614. [Google Scholar] [CrossRef] [Green Version]
  190. Riboulet-Bisson, E.; Sturme, M.H.; Jeffery, I.B.; O’Donnell, M.M.; Neville, B.A.; Forde, B.M.; Claesson, M.J.; Harris, H.; Gardiner, G.E.; Casey, P.G.; et al. Effect of Lactobacillus salivarius bacteriocin Abp118 on the mouse and pig intestinal microbiota. PLoS ONE 2012, 7, e31113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Murphy, E.F.; Cotter, P.D.; Hogan, A.; O’Sullivan, O.; Joyce, A.; Fouhy, F.; Clarke, S.F.; Marques, T.M.; O’Toole, P.W.; Stanton, C.; et al. Divergent metabolic outcomes arising from targeted manipulation of the gut microbiota in diet-induced obesity. Gut 2013, 62, 220–226. [Google Scholar] [CrossRef] [PubMed]
  192. Clarke, S.F.; Murphy, E.F.; O’Sullivan, O.; Ross, R.P.; O’Toole, P.W.; Shanahan, F.; Cotter, P.D. Targeting the microbiota to address diet-induced obesity: A time dependent challenge. PLoS ONE 2013, 8, e65790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Mills, S.; Stanton, C.; Hill, C.; Ross, R.P. New developments and applications of bacteriocins and peptides in foods. Annu. Rev. Food Sci. Technol. 2011, 2, 299–329. [Google Scholar] [CrossRef] [PubMed]
  194. Zhou, H.; Xie, Y.; Liu, H.; Jin, J.; Duan, H.; Zhang, H. Effects of two application methods of plantaricin BM-1 on control of Listeria monocytogenes and background spoilage bacteria in sliced vacuum-packaged cooked ham stored at 4 °C. J. Food Prot. 2015, 78, 1835–1841. [Google Scholar] [CrossRef] [PubMed]
  195. Chandrakasan, G.; Rodrıguez-Hernandez, A.-I.; del Rocıo Lopez-Cuellar, M.; Palma-Rodriquez, H.-M.; Chavarria-Hernandez, N. Bacteriocin encapsulation for food and pharmaceutical applications: Advances in the past 20 years. Biotechnol. Lett. 2019, 41, 453–469. [Google Scholar] [CrossRef]
  196. Udompijitkul, P.; Paredes-Sabja, D.; Sarker, M.R. Inhibitory effects of nisin against Clostridium perfringens food poisoning and nonfood-borne isolates. J. Food Sci. 2012, 77, M51–M56. [Google Scholar] [CrossRef]
  197. Bartoloni, A.; Mantella, A.; Goldstein, B.P.; Dei, R.; Benedetti, M.; Sbaragli, S.; Paradisi, F. In-vitro activity of nisin against clinical isolates of Clostridium difficile. J. Chemother. 2004, 16, 119–121. [Google Scholar] [CrossRef]
Figure 1. Keystone points in evaluation of potential application of bacteriocins as powerful antimicrobials in biopreservation of food commodities and/or new generation of drugs.
Figure 1. Keystone points in evaluation of potential application of bacteriocins as powerful antimicrobials in biopreservation of food commodities and/or new generation of drugs.
Foods 11 03145 g001
Table 1. Examples of some proposed food and feed-related applications for some of recently reported bacteriocins.
Table 1. Examples of some proposed food and feed-related applications for some of recently reported bacteriocins.
BacteriocinProducerArea of Proposed ApplicationReference
Bovicin HC5 and nisinStreptococcus bovis HC5Control of Alicyclobacillus acidoterrestris in fruit juices[109]
Bovicin HC5Streptococcus bovis HC5Clostridium tyrobutyricum, a pathogen associated with spoiled mango pulp[110]
Enterococin AEnterococcus faecium MMRAControl of Listeria monocytogenes in sliced dry-cured ham[111]
Bacteriocin RSQ04Lactococcus lactis CGMCC20699Evaluation of activity against Listeria monocytogenes in model food system[112]
Pyocin QDD1Pseudomonas aeruginosa QDD1Biocontrol of foodborne pathogens Staphylococcus aureus and Bacillus cereus[113]
Bacteriocin LSB1Lactiplantibacillus plantarum LSB1Activity against Staphylococcus argenteus planktonic cells and biofilm[114]
Bacteriocins ST20Kc and ST41KcEnterococcus faecium ST20Kc and ST41KcControl of Listeria monocytogenes and vancomycin-resistant entorococci[115]
Bacteriocin LSX01 Lacticaseibacillus paracasei LSX01 Reduction of planktonic cells of Staphylococcus aureus[116]
Bacteriocin CTC494Latilactobacillus sakei CTC494Anti-listerial activity in vacuum packaged cooked ham[117]
Six bacteriocinsLacticaseibacillus casei and Lactiplantibacillus plantarumActivity against 8 different Listeria monocytogenes strains [118]
Bacteriocins ST651ea, ST7119ea, and ST7319eaEnterococcus faecium ST651ea, ST7119ea, and ST7319eaControl of Listeria monocytogenes and vancomycin-resistant enterococci in GIT model system[119]
Bacteriocin Sak-59Lactobacillus sakei B-RKM 0559Activity against meat spoilage bacteria strains of Listeria monocytogenes, Staphylococcus aureus, and pathogenic strains of Serratia marcescens and Escherichia coli[120]
Lactocin 63Loigolactobacillus coryniformis FZU63Antimicrobial mode of action against Shewanella putrefaciens[121]
Bacteriocin R23Lactiplantibacillus plantarum R23Anti-Listeria monocytogenes activity[122]
Enterocin LD3 and Plantaricin LD4Enterococcus faecium LD3 and Lactiplantibacillus plantarum LD4Synergistic effect against Staphylococcus aureus subsp. aureus ATCC25923, Salmonella enterica subsp. enterica serovar Typhimurium ATCC13311, Proteus mirabilis ATCC43071, Pseudomonas aeruginosa ATCC27853, and Escherichia coli ATCC25922[123]
Bacteriocins ST1607V, ST2104V and ST3105VPediococcus acidilactici ST1607V, ST2104V and ST3105VBactericidal mode of action against Listeria monocytogenes ATCC7644 and Enterococcus faecium ATCC19434[48]
Bacteriocin BM1829Companilactobacillus crustorum MN047Reduction of Escherichia coli and Staphylococcus aureus[124]
Enterocin K2B1Enterococcus faecalis K2B1Control of foodborne pathogens in dairy products[125]
Bacteriocin OS1Enterococcus hirae OS1Anti-Listeria activity[126]
Bacteriocins ST3522BG and ST3633BGPediococcus acidilactici ST3522BG and Pedioccocus pentosaceus ST3633BGAnti-Listeria activity in silage fermentation models system[127]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Todorov, S.D.; Popov, I.; Weeks, R.; Chikindas, M.L. Use of Bacteriocins and Bacteriocinogenic Beneficial Organisms in Food Products: Benefits, Challenges, Concerns. Foods 2022, 11, 3145. https://doi.org/10.3390/foods11193145

AMA Style

Todorov SD, Popov I, Weeks R, Chikindas ML. Use of Bacteriocins and Bacteriocinogenic Beneficial Organisms in Food Products: Benefits, Challenges, Concerns. Foods. 2022; 11(19):3145. https://doi.org/10.3390/foods11193145

Chicago/Turabian Style

Todorov, Svetoslav Dimitrov, Igor Popov, Richard Weeks, and Michael Leonidas Chikindas. 2022. "Use of Bacteriocins and Bacteriocinogenic Beneficial Organisms in Food Products: Benefits, Challenges, Concerns" Foods 11, no. 19: 3145. https://doi.org/10.3390/foods11193145

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop