Next Article in Journal
FDA-Approved Small Molecule Compounds as Drugs for Solid Cancers from Early 2011 to the End of 2021
Next Article in Special Issue
Efficient Bioconversion of Stevioside and Rebaudioside A to Glucosylated Steviol Glycosides Using an Alkalihalobacillus oshimesis-Derived Cyclodextrin Glucanotransferase
Previous Article in Journal
Diterpenoid Alkaloids Isolated from Delphinium brunonianum and Their Inhibitory Effects on Hepatocytes Lipid Accumulation
Previous Article in Special Issue
Changes in Organic Acids, Phenolic Compounds, and Antioxidant Activities of Lemon Juice Fermented by Issatchenkia terricola
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enterocins: Classification, Synthesis, Antibacterial Mechanisms and Food Applications

1
Department of Food Science and Nutrition, Zhejiang University, Hangzhou 310058, China
2
College of Agriculture and Forestry, Linyi University, Linyi 276005, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(7), 2258; https://doi.org/10.3390/molecules27072258
Submission received: 27 February 2022 / Revised: 23 March 2022 / Accepted: 28 March 2022 / Published: 30 March 2022

Abstract

:
Enterococci, a type of lactic acid bacteria, are widely distributed in various environments and are part of the normal flora in the intestinal tract of humans and animals. Although enterococci have gradually evolved pathogenic strains causing nosocomial infections in recent years, the non-pathogenic strains have still been widely used as probiotics and feed additives. Enterococcus can produce enterocin, which are bacteriocins considered as ribosomal peptides that kill or inhibit the growth of other microorganisms. This paper reviews the classification, synthesis, antibacterial mechanisms and applications of enterocins, and discusses the prospects for future research.

1. Introduction

Enterococci are catalase-negative, gram-positive cocci that have an average GC content of less than 40% and produce lactic acid as the main end product of glucose fermentation. Enterococcus is a type of facultative anaerobic bacteria that can tolerate temperatures of 10–45 °C and grow best at 35 °C [1]. Enterococci are widely distributed in the natural environment and are part of the normal flora in the intestinal tract of humans and animals. However, in recent years, the bacteria has gradually evolved into a pathogenic and multi-drug resistant branch, and has become an important pathogen that threatens the health of human beings, especially hospitalized patients [2,3]. The drug-resistant E. faecium is a member of the so-called “ESKAPE” classification (E. faecium, Staphylococcus aureus, Klebsiella pneumoniae, Pseudomonas aeruginosa and Enterobacter spp.), which is listed by the Infectious Diseases Society of America (IDSA) as an essential factor in global medical treatment difficulty and death [4].
Previous studies have shown that pathogenic drug-resistant bacteria were significantly more prevalent in hospitalized patients, doctors and nurses than healthy community populations. The major cause of infection with drug-resistant enterococci is its colonization and unbalanced proliferation in the intestine [5]. In recent years, multidrug-resistant enterococci isolated from animal samples have become increasingly frequent [6,7,8], and animal-derived drug-resistant enterococci have higher pathogenic potential and transmission threat than healthy population isolates [9,10]. Farmed animals and pets, as hosts of Enterococcus, play an important role in the accumulation of evolutionary drug resistance and pathogenic factors and the transmission between humans and animals. At the same time, intensive and large-scale farming methods have emerged because traditional animal production methods can no longer meet the growing human demand for animal food such as meat, eggs and milk. Under this situation, the immunity of animals is weakened due to the excessive breeding density, and the abuse of antibiotics leads to a great increase in the probability of animals carrying pathogenic and drug-resistant Enterococcus, making farmed animals become an important carrier and source of the pathogenic bacteria. Therefore, the use of non-antibiotic methods to specifically reduce or eliminate pathogenic drug-resistant Enterococcus carried in the host, while avoiding the development of new drug resistance, will be an ideal method to prevent and treat the disease.
Unlike other pathogenic bacteria, certain non-pathogenic Enterococcus isolates are also used as food probiotics or clinical therapeutic probiotics for humans and pets [11,12], and as feed additives for disease control and growth promotion in the breeding industry [13]. This is because many strains of Enterococcus can produce bacteriocins, which are known as enterocins.
Bacteriocins are peptide antibiotics secreted by Enterococcus during the metabolic process. They are extracellular secreted proteins with certain bacteriostatic activities, and are usually encoded by plasmid genes or chromosomal genes and synthesized by ribosomes [14]. Most bacteriocins only have damaging effects on closely related bacteria, but there are still parts of them with inhibitory effects that are not limited to homologous bacteria. The bacteriocin-producing strain is immune to the bacteriocins it produces [15]. Bacteriocin constitutes a special defense mechanism of bacteria, which can inhibit the growth of other bacteria and enable the bacteriocin-producing bacteria to gain a competitive advantage, which is of great significance for maintaining the balance of flora in the ecological environment.
Bacteriocin can inhibit the food-borne pathogen Listeria monocytogenes, and has been applied in food processing to improve food quality and safety, which has produced huge economic and social benefits [16]. At present, nisin is the most widely researched and applied bacteriocin. Nisin was first introduced to the market in the United Kingdom in 1953, and more than 50 countries have successively approved the application of nisin. In 1969, the Expert Committee of the Food and Agriculture Organization of the United Nations and the World Health Organization unanimously approved the use of nisin as a safe additive in food [17]. The successful development of nisin into a commercial product model greatly stimulated research on bacteriocin. In the future, enterocins may bring huge application value and economic benefits in food, feed addition or other aspects. This review will provide readers with important information on enterocins in terms of their biological properties, structure–function relationships, antibacterial mechanisms and applications in food systems, which will help in mining effective enterocins for various pathogenic bacteria and exploring the intervention method of “bacteria treatment with bacteria”.

2. Classification of Enterocins

In 1993, Klaenhammer proposed the first classification method of lactobacillus bacteriocin [18]. With the complexity of the bacteriocin structure and the continuous expansion of the bacteriocin family, the classification scheme has been continuously revised. At present, the latest and most authoritative three-category classification method has replaced the previous four-category classification scheme [19]. The classification standard of this classification method is based on the main structure, molecular weight, post-transcriptional modification and other biosynthetic mechanisms and biological activities of bacteriocin.
Class I bacteriocins refer to small-molecule (<10 kDa) thermostable bacteriocins with extensive post-translational modifications. Enzymatic modifications during biosynthesis, such as dehydration to lanthionine, glycosylation, head-to-tail cyclization and others [20], provide such bacteriocins with uncommon amino acid structures or structures that affect their activity. Nisin is a representative of class I bacteriocins. Cytolysin and enterocin AS-48 [21] are common class I enterocins. Cytolysin, which belongs to class Ia (lanthipeptides), are encoded by hemolytic virulence factors, so they do not have potential for food and drug applications [22]. Enterocin AS-48 was the first circular enterocin that was described and then classified into class Ib. Its N-terminal and C-terminal are bound by a peptide bond formed between the N-terminal methionine and the C-terminal tryptophan. Enterocin AS-48 has a total of 70 amino acid residues, containing no modified amino acid residues or disulfide bonds [23]. The secondary and tertiary structure of enterocin AS-48 has been reported. The structure of enterocin AS-48 consists of a spherical array of five α-helices surrounding a compact hydrophobic core [16]. The round shape and compact creases provide significant stability to the enterocin AS-48 molecule against extreme pH, heat and denaturants. Therefore, enterocin AS-48 has been widely applied to food systems due to its stability. Enterocin F4–9 belongs to class Ie (glycocins), containing glycosylated amino acid residues. Enterocin F4–9 does not completely kill the indicator bacteria but inhibits their growth [24], and the mechanism of action is unclear.
Class II bacteriocins refer to unmodified small-molecule (<10 kDa) thermostable bacteriocins after translation, which do not involve the formation of uncommon amino acids. Therefore, they do not require special enzymes other than signal peptides or transporters to complete the maturation and activation of such bacteriocins. Most enterocins are class II and these bacteriocins can inhibit the growth of L. monocytogenes [25]. Furthermore, most of them contain a double glycine signal peptide sequence between amino acid residues 15 and 30 of the N-terminal [26]. Enterocin A was a representative bacteriocin of enterocins and was the first enterocins of class IIa to be isolated, purified and relatively well-studied. The presence of the entA gene was detected in the first E. faecium strain E. faecium DO to complete whole-genome sequencing, but it did not exhibit bacteriocin activity. Enterocin X is a class IIb (two-peptide) bacteriocin derived from E. faecium KU-B5, and is composed of two antibacterial peptides (Xα and Xβ). Xα and Xβ are two peptides with different properties, which can produce antibacterial activity when they are mixed in equal proportions [27]. Class IIc enterocins (a type of N-terminal leaderless peptide) are composed of homologous peptides, encoded by tandemly repeated genes, and these peptides have a high homology of more than 70% in amino acid sequences. Enterocin L50 is a plasmid-encoded broad spectrum bacteriocin composed of enterocin L50A and enterocin L50B. These peptides share 72% sequence identity and are composed of 44 and 43 amino acids, respectively. Both of the two enterocins have antibacterial activity on their own, with entL50A being the most active. In addition, significant synergistic effects were observed when the two enterocins were used in combination, especially with specific indicator strains [28]. Enterocin B belongs to class IId and shows strong homology to carnobacteriocin A [29].
Class III bacteriocins include unmodified macromolecular heat-labile bacteriocins with molecular weights greater than 10 kDa. They have bacteriolytic activity and are the most poorly-characterized class of bacteriocins so far. Enterolysin A belongs to class III and has a bacteriolytic mode of action. Mature Enterolysin A is composed of 316 amino acids with a calculated molecular weight of 34,501 [30].
Although the taxonomic status of the genus Enterococcus was only established in 1982, many Enterococcus bacteriocins have been isolated and purified in recent years, and most of them were obtained from Enterococcus faecalis and Enterococcus faecium. According to the latest classification standard of bacteriocins, the main categories and properties of enterococcal bacteriocins are listed in Table 1.

3. Synthesis of Enterocins

Enterocins, produced by ribosome, have multiple functions in the innate immunity of bacteria and in microbial interactions [38]. In general, gram-negative bacteriocins are large-domain protein toxins with receptor-mediated narrow spectrum activity, while gram-positive bacteriocins are predominantly peptides [39]. Therefore, most enterocins are plain peptides, which may or may not contain disulfide bonds and cysteines. Genes that produce active enterocins are usually located in operon clusters, including plasmids, transposons and chromosomes [20]. The expression of these operons is induced by the auto-inducer peptide and regulated by a two-component control system [40] or a three-component control system [41].
Different classes of enterocins have different biosynthetic pathways (Figure 1) [42]. For class I lanthipeptides, the serine and cysteine residues in their propeptides undergo extensive post-translational modification. The lanthipeptides biosynthesis operons’ clusters contain genes including LanA (structural gene), LanB and LanC (encoding enzymes required for modification), LanT (encoding transport proteins), LanP (encoding a serine proteinase that is capable of removing the leader sequence), LanI and LanEFG (immunity gene). Their biosynthetic pathway follows the general scheme shown below: (1) formation of propeptides; (2) modification reactions; (3) proteolytic activation of the leader peptide; and (4) translocation of the mature or modified propeptides across the cytoplasmic membrane. Class II enterocins, however, do not undergo extensive post-translational modification. Propeptides are removed and exported out from the cell after the formation.
Once the enterocin is synthesized, it exists in an inactive form containing a leader peptide at the N-terminus. The leader sequence is an important factor to maintain the inactivity of the enterocins in the producing strain and can promote the interaction between the bacteriocin and the transporter. Gene clusters encoding leader peptides generally include structural, regulatory, transport and related modified genes. The leader peptide chain is usually cleaved through the transport system or the general secretion system of the cell. As more enterocins have been discovered and studied, several enterocins have been confirmed to be N-terminal leaderless peptides, such as enterocin L50, enterocin Q, enterocin EJ97 and enterocin RJ-11. These enterocins are divided into class IIc, as mentioned above.

4. Production and Purification of Enterocins

Enterocin genes usually have a low expression in wild strains. To improve the production of enterocins for better utilization, many researchers have investigated the production of recombinant proteins expressed by heterologous enterocins. Currently, three different heterologous expression methods have been proven to be effective: E. coli, yeast and Lactobacillus [44]. We sum up some common enterocins that have been heterologously expressed in three different systems (Table 2). Generally, heterologous expression in the above systems is efficient because these strains have relatively clear and mature genetic backgrounds that are convenient in terms of controlling the gene expression and achieving low culture costs with higher production of proteins of interest. Besides, chemical synthesis is also a great alternative way to attain pure enterocins without the complex purification of microbial fermentation.
The purification of enterocins is critical for the further study and application of enterocin after the enterocin has been produced. Purification methods can be divided into three categories according to the quality of the purified enterocins, which are crude extraction, moderate extraction and high-purity extraction. The crude extraction methods include ammonium sulfate precipitation, organic solvent extraction and adsorption. This method is suitable for the extraction of enterocin with unknown physical and chemical properties. The moderate purification methods can remove the impurity proteins in the crude enterocins, and include gel-filtration chromatography, ion-exchange chromatography and hydrophobic-interaction chromatography. Besides, high-purity extraction that involving reversed-phase HPLC and electrophoresis can produce enterocins with a purity over 95%, which meets the requirements of mass spectrometry and amino acid sequencing.

5. Antibacterial Activity and Mechanisms of Action

Enterocins are polypeptides or precursor polypeptides with bacteriostatic activity produced by enterococcus during metabolism. Previous studies have shown that enterocins exhibit strong antimicrobial activity against important foodborne pathogens, including Staphylococcus aureus, Listeria monocytogenes and Salmonella enterica et al. [53]. For example, purified enterocin E-760 has been proven to inhibit the growth of 24 species of gram-negative bacteria and 3 species of gram-positive bacteria. The MIC of enterocin E-760 for these indicator bacteria ranges from 0.1 μg/mL to 3.2 μg/mL [57].
The antibacterial mechanism of bacteriocins, which differs from the antibiotics, is closely connected with the classification of bacteriocins and the indicator bacteria [58]. It is crucial for the normal cell to maintain the integrity of cell walls and membranes. Bacteriocin inhibits other bacteria by perforating on target cells, inhibiting peptidoglycan synthesis, inhibiting protein synthesis by interacting with ribosomes or tRNA, and directly degrading target cell DNA [59,60].
The antibacterial mechanisms of enterocins can be classified into the cell envelope and inhibition of gene expression (Figure 2). Several enterocins can interact with specific bacterial cell receptors, including lipid II and other related cell wall precursors, undecaprenyl pyrophosphate phosphatase, the mannose phosphotransferase system, the maltose ABC transporter or a Zn-dependent metallopeptidase [61]. In addition, most class II enterocins can induce membrane permeabilization and then cause the leakage of target cells. Enterocin P is an amphipathic and cationic peptide, which was proven to interact with the negatively charged bacterial membranes of the target cells and then cause death [62]. The majority of class III enterocins, which are large and heat-labile proteins, can break down the sensitive cells by catalyzing cell-wall hydrolysis. Enterolysin A can hydrolyze peptide bonds in the peptidoglycan of target cells [30]. Besides, enterocins can control their target bacteria by interfering in DNA replication and mRNA synthesis and transcription [63].

6. Application of Enterocins in Food Industry

Many enterocins have been found to be active against pathogens and used as food preservatives for food protection. Examples of applications of enterocins in three main food systems are listed in Table 3.

6.1. Dairy Products

Milk is a highly nutritious food and provides a perfect environment for a large variety of microorganisms [64]. Dairy products are easily contaminated by various pathogenic microorganisms in the process of production, packaging, transportation and sales, which is considered to be a great threat to the quality and safety of food. Staphylococcus aureus and Listeria monocytogenes are common foodborne pathogens and have caused food poisoning outbreaks many times [65]. Enterocins in dairy products can naturally prevent the growth of microorganisms.
Enterocin AS-48, produced by E. faecalis A-48-32, is a circular bacteriocin with a wide-range bactericidal activity against various gram-positive bacteria, such as S. aureus, Bacillus cereus and L. monocytogenes [23,66]. The effectiveness of enterocin AS-48, which controls S. aureus and B. cereus in skim milk and fresh cheese, has been determined. Enterocin AS-48 combined with moderate heat treatment (65 °C, 5 min) showed a strong anti-staphylococcal synergy at a concentration of 20 μg/mL, and was able to remove staphylococci after six hours of incubation [67]. In another research study, Arantxa tested enterocin AS-48 against B. cereus in milk and cheese. The results of the experiments revealed that no viable B. cereus cells were found after 72 h incubation in food systems that were inoculated with B. cereus and produced the strain of enterocin AS-48.
In addition, Lauková with his lab members were the first to detect enterocin CCM 4231 produced by E. faecium CCM 4231, isolated from the rumen content of a calf [68]. They carried out tests for the antibacterial activity of this enterocin against different pathogens especially in dairy products. In skim milk, live cells of the S. aureus decreased from 1010 cfu/mL to 102 cfu/mL 24 h after the addition of enterocin CCM 4231, but the control did not show a decrease in staphylococcal count (enterocin was not added). During the preparation of yogurt, significant inhibition of the S. aureus was observed after the addition of enterocin CCM 4231 for 3.5 h [69]. Besides, inhibition of enterocin CCM 4231 (3200 AU/mL) was found in experimental cheese samples via the addition of Listeria strains during the preparation of St. Paulin’s cheese, from an initial concentration of 6.7 ± 0.06 log cfu/g to 1.9 ± 0.01 log cfu/g after one week, compared to a control treatment with pathogen levels of 6.0 ± 0.01 log cfu/g after one week [70]. Similarly, in other studies, many enterocins were also proven to be inhibitory to pathogens, including enterocin EJ97 [35], enterocin E-760 [57] and enterocin CRL 35 [71]. These above results clearly demonstrate the antibacterial potential of enterocins, which have great application prospects in the processing of dairy products.

6.2. Meat Products

In addition to dairy products, other products of animal origin, namely meat and fish, are also very beneficial to the growth of microorganisms due to their moisture and rich nutrients. Foods of animal origin are the major hosts for many foodborne pathogens, including Salmonella, Campylobacter, L. monocytogenes and the Shiga toxin-producing strains of Escherichia coli [72]. According to the data collected by the European Food Safety Authority (EFSA), a total of 5196 weak and strong outbreaks involving meat products were reported in Europe alone in 2013. In these outbreaks, Salmonella enterica remained the most common pathogen (22.5%), followed by viruses (18.1%) and bacterial toxins (16.1%) [73]. The potential for zoonotic diseases and their ability to produce enough toxins to cause illness and even death illustrates the seriousness of the situation. Food poisoning causes millions of sporadic illnesses, chronic complications, and even large-scale and difficult outbreaks across many countries and nations. Therefore, the use of enterocins has potential in the preservation of meat.
The anti-listeria effect of enterocins A and B, produced by E. faecium CTC492, has been demonstrated in meat and meat products (cooked ham, chicken breast and fermented sausages). Enterocins A and B (4800 AU/g) diminished the number of Listeria by 7.98 log cycles in cooked ham. In chicken breasts stored at 7 °C for 7 days, 4800 AU/cm2 enterocins A and B reduced the number of Listeria by 5.26 log cycles compared to the control [74]. After that, a number of researchers combined the application of enterocins A and B and high hydrostatic pressure (HHP) to extend the shelf-life of food products. Enterocins were applied to cooked and dry cured ham with L. monocytogenes and Salmonella and subjected to high-pressure treatment at 600 MPa. Salmonella and L. monocytogenes were not detected at 25 g in both cooked and dried prosciutto and remained at this level during the entire storage [75].
The potential of enterocin AS-48 to inhibit food-derived pathogens in meat products has been demonstrated. No viable L. monocytogenes was detected after culturing at 20 °C for 6 and 9 days in a meat sausage model system supplemented with enterocin AS-48 [76]. In the cooked ham model system, enterocin AS-48 alone (20, 40, and 60 μg/g) was active against L. monocytogenes at 5 and 15 °C, but it could not avoid the regrowth of Listeria during the 60 days of shelf-life. However, enterocin AS-48 combined with chemical preservatives or heat can improve the efficacy of anti-Listeria during storage at 5 °C. The most effective combination was AS-48-nitrite/nitrate (0.007%), which caused Listeria to fall below detection levels from the start to the end of storage [77]. As with the enterocins A and B, enterocin AS-48 combined with HHP can inhibit the growth of S. enterica, L. monocytogenes and S. aureus. In another research study, 148 AU/g enterocin AS-48 caused a dramatic reduction of 5.5 log cfu/g in L. monocytogenes and a significant inhibition for Salmonella at the end of ripening [78]. Enterocin LM-2 was also combined with HHP to suppress the growth of Salmonella and L. monocytogenes [79]. Besides, enterocin CCM 4231 was reported to be applied to inhibit L. monocytogenes in dry fermented Hornád salami, resulting in a 1.67 log cycle reduction in L. monocytogenes [80].

6.3. Fruits and Vegetables Products

Fresh fruits and vegetables are an important part of the human diet, and public health agencies recommend eating at least five servings of fruits and vegetables daily. These fruits and vegetables are rich in minerals and vitamins, along with trace elements such copper, manganese and zinc, which are components of various enzyme cofactors required for different metabolic reactions. There has been a growing demand for fresh and healthy food, especially fruits and vegetables, during the last decade. However, the consumption of fruit and vegetable products is also susceptible to contamination by pathogens. Since vegetables and fruits are organisms with different compositional properties, their microbial fauna can vary significantly depending on the pH, water activity and availability of nutrients [81]. Mesophilic enteric pathogens such as E. coli, Salmonella and L. monocytogenes are common pathogens that cause disease in humans. Salmonella and E. coli O157: H7 can cause infection even at low doses and grow very rapidly [82]. Enterocins have been investigated in fruits and vegetables in relation to the preservation of such products.
Plant-derived foods have been studied in detail as targets for the application of enterocin AS-48. When enterocin AS-48 (6 μg/mL) is added to low-acid canned vegetables (tomato paste, canned peach syrup and canned pineapple juice), Bacillus coagulans cells are completely or partially inactivated [83]. Enterocin AS-48 is effective in inactivating L. monocytogenes, B. cereus and Bacillus weihenstephanensis in sprouts, as well as L. monocytogenes in sliced fruits and whole fruit pieces when used as washing treatment alone or in combination with other chemical antimicrobials [84,85,86]. Enterocin AS-48 can be a good candidate to apply in the preservation of fruits juices because of the strong stability of this enterocin in juices [87]. When it was added to juices contaminated with endospores of Alicyclobacillus acidoterrestris with low concentrations, the bacteria could be inhibited completely for up to two weeks in fresh apple and orange juices, and for up to two to three months in various commercial fruit juices under storage temperatures [88]. Widespread enterocin AS-48 was tested for biopreservation against aerobic mesophilic endospore-forming bacteria in ready-to-eat vegetable foods (soups and purees). B. cereus LWL1, in six vegetable products tested (natural vegetable cream, asparagus cream, traditional soup and homemade traditional soup), was inhibited totally by adding AS-48 (10 μg/mL) for up to 30 days at 6, 15 and 22 °C [89].
As with the production of other food products, HHP was combined with enterocin AS-48 in the fruits and vegetables. As is shown in previous research studies, the combination of AS-48 and HHP was more effective in maintaining the quality of cherimoya pulp than HHP treatment alone during storage [90].
Enterocin 416K1, produced by E. casseliflavus IM 416K1, demonstrated strong anti-listerial activity with artificially contaminated L. monocytogenes ready-to-eat vegetables and fruits [91]. The research showed that in apples and grapes, the number of viable Listeria in the control dropped to about 2 log cfu/g within one day and remained constant until the experiment ended. In the treated sample, L. monocytogenes was completely inhibited. This was achieved within the first hour of using chitosan and enterocin together, within 8 h for the enterocin-only sample, and within 72 h for the chitosan-only sample. Besides, enterocin EJ97 can control the vegetative cells and spores of Geobacillus stearothermophilus, which causes sour spoilage of high-water-activity and low-acid canned foods [92].
Most enterocins show a certain degree of bacteriostatic activity. However, a number of enterocins only show strong bacteriostatic activity when they act synergistically with other substances. Enterocin KT2W2G, combined with cinnamon oil at 4:6 ratios, exhibited the strongest synergistic effect spoilage microorganisms (Lactococcus lactis, E. faecalis, Klebsiella pneumonia, Serratia marcescens pichia and Candida metapsilosis), whereas enterocin KT2W2G displayed no activity when using alone [93].
Table 3. Application of enterocins in food systems.
Table 3. Application of enterocins in food systems.
Food SystemsEnterocins Producing StrainAdditional TreatmentTarget OrganismsReferences
Dairy productsEnterocin AS-48E. faecalis A-48-32moderate heat (65 °C, 5 min); Cocultivation;Staphylococcus aureus;
Bacillus cereus
[23,92]
Enterocin CCM 4231E. faecium CCM 4231add purified enterocin CCM 4231 (concentration 3200 AU/mL) Escherichia coli;
Listeria monocytogenes;
Staphylococcus aureus
[68,78]
Enterocin EJ97E. faecalis EJ97sodium nitrite, sodium benzoate, et.alListeria monocytogenes;[35]
Enterocin E-760E. faecium NRRL B-30745 Salmonella enterica;
Yersinia enterocolitica;
[54]
Dairy productsEnterocin CRL 35E. faecium CRL 35 Listeria monocytogenes;[69]
Meat productsEnterocins A and BE. faecium CTC492 Salmonella entericai; Listeria monocytogenes; Staphylococcus aureus;[72,93]
Enterocin AS-48E. faecalis A-48-32added at concentrations of 30 or 40 μg/g;
high hydrostatic pressure
Staphylococcus aureus; Salmonella enterica;
Listeria monocytogenes;
[76,94]
Enterocin CCM 4231E. faecium CCM 4231 Listeria monocytogenes;[78]
Enterocin LM-2E. faecium LM-2high hydrostatic pressureSalmonella enterica;
Listeria monocytogenes;
[77]
Fruits and vegetables productsEnterocin AS-48E. faecalis A-48-32Washing with chemical preservatives;
heat (80–95 °C for 5 min); nisin and phenolic compounds; cinnamic and hydrocinnamic acids, etc
Bacillus cereus;
Listeria monocytogenes;
Staphylococcus aureus; Bacillus coagulans; Paenibacillus sp.; Bacillus weihenstephanensis; Pantoea; Leuconostocs
[81,83,86,88,95]
Enterocin 416K1E. casseliflavus IM 416K1chitosanListeria monocytogenes[89]
Fruits and vegetables productsEnterocin EJ97E. faecalis EJ97sodium tripolyphosphate and sodium nitriteBacillus coagulans[96]
Enterocin KT2W2GE. faecalis KT2W2Gessential oils; cinnamon oilLactococcus lactis subsp. Lactis; Enterococcus faecalis; Klebsiella pneumonia; Klebsiella variicola; Serratia marcescens[91]

7. Conclusions

In recent years, there has been a growing interest in the use of bacteriocins to partially replace traditional antibiotics in humans or animals. Bacteriocin has the characteristics of high efficiency, high specificity, low toxicity, less drug-resistance, etc. It meets food safety grade requirements and is widely used in the food industry to improve food quality and safety, which has produced huge economic and social benefits. So far, nisin is the only bacteriocin approved for use as a food preservative in more than 50 countries. Various tests have proved that nisin has extremely high safety. Its median lethal dose (LD50) to mice is about 7000 mg/kg, which is close to the LD50 of ordinary table salt. Enterocins, as a bacteriocin, have very broad prospects in food and medical applications.
Although many patents on enterocins have been reported [94,95,96,97], these enterocins still have limitations. The moderate antimicrobial effects and limited specificity are major limitations in the application of enterocins in the food industry [98]. Additionally, certain enterocins may be inactivated by food components, which is an important problem and it is almost impossible to find the exact cause of the inhibition [87]. It has been nearly a century since the discovery and utilization of bacteriocins; however, the research on enterocins is still mainly limited to the aspects of mining and application. Although several mechanisms of action have been reported, they mainly focus on the target and regulation of action. There is still a lack of deeper and more comprehensive research, and an in-depth understanding of the classification, expression and mechanisms would be beneficial in terms of the large scale production and wide application of enterocins.

Author Contributions

Investigation, Y.W. (Yajing Wu); resources, X.P. and Y.W. (Yansha Wu); data curation, Y.W. (Yajing Wu) and X.L.; writing—original draft preparation, Y.W. (Yajing Wu); writing—review and editing, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Project of Shandong Province Higher Educational Outstanding Youth Innovation Team (2019KJF011), the Natural Science Foundation of Shandong Province, China (ZR2019ZD21), the Taishan Scholars Program of Shandong Province, China (ts20190955), and the National Key Research and Development Program of China (2019YFE0103900).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable. No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Byappanahalli, M.N.; Nevers, M.B.; Korajkic, A.; Staley, Z.R.; Harwood, V.J. Enterococci in the Environment. Microbiol. Mol. Biol. 2012, 76, 685–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gilmore, M.S.; Lebreton, F.; van Schaik, W. Genomic transition of enterococci from gut commensals to leading causes of multidrug-resistant hospital infection in the antibiotic era. In Current Opinion in Microbiology; Elsevier Current Trends: Cambridge, MA, USA; Volume 16, pp. 10–16.
  3. Prieto, A.M.G.; van Schaik, W.; Rogers, M.R.; Coque, T.M.; Baquero, F.; Corander, J.; Willems, R.J. Citation: Global Emergence and Dissemination of Enterococci as Nosocomial Pathogens: Attack of the Clones? Front. Microbiol. 2016, 7, 788. [Google Scholar]
  4. Boucher, H.W.; Talbot, G.H.; Bradley, J.S.; Edwards, J.E.; Gilbert, D.; Rice, L.B.; Scheld, M.; Spellberg, B.; Bartlett, J. Bad Bugs, No Drugs: No ESKAPE! An Update from the Infectious Diseases Society of America. Clin. Infect. Dis. 2009, 48, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Arias, C.A.; Murray, B.E. The rise of the Enterococcus: Beyond vancomycin resistance. Nat. Publ. Group 2012, 10, 266–278. [Google Scholar] [CrossRef] [Green Version]
  6. Iweriebor, B.C.; Obi, L.C.; Okoh, A.I. Virulence and antimicrobial resistance factors of Enterococcuss isolated from fecal samples from piggery farms in Eastern Cape, South Africa Ecological and evolutionary microbiology. BMC Microbiol. 2015, 15, 136. [Google Scholar] [CrossRef] [Green Version]
  7. Tamang, M.D.; Moon, D.C.; Kim, S.-R.; Kang, H.Y.; Lee, K.; Nam, H.-M.; Jang, G.-C.; Lee, H.-S.; Jung, S.-C.; Lim, S.-K. Detection of novel oxazolidinone and phenicol resistance gene optrA in enterococcal isolates from food animals and animal carcasses. Vet. Microbiol. 2017, 201, 252–256. [Google Scholar] [CrossRef]
  8. Wang, Y.; Lv, Y.; Cai, J.; Schwarz, S.; Cui, L.; Hu, Z.; Zhang, R.; Li, J.; Zhao, Q.; He, T.; et al. A novel gene, optrA, that confers transferable resistance to oxazolidinones and phenicols and its presence in Enterococcus faecalis and Enterococcus faecium of human and animal origin. J. Antimicrob. Chemother. 2015, 70, 2182–2190. [Google Scholar] [CrossRef] [Green Version]
  9. Abdel-Moein, K.A.; El-Hariri, M.D.; Wasfy, M.O.; Samir, A. Occurrence of ampicillin-resistant Enterococcus faecium carrying esp gene in pet animals: An upcoming threat for pet lovers. J. Glob. Antimicrob. Resist. 2017, 9, 115–117. [Google Scholar] [CrossRef]
  10. Van den Bunt, G.; Top, J.; Hordijk, J.; de Greeff, S.C.; Mughini-Gras, L.; Corander, J.; van Pelt, W.; Bonten, M.J.; Fluit, A.C.; Willems, R.J. Intestinal carriage of ampicillin-and vancomycin-resistant Enterococcus faecium in humans, dogs and cats in the Netherlands. Antimicrob. Chemother. 2018, 73, 607–614. [Google Scholar] [CrossRef]
  11. Bybee, S.; Scorza, A.; Lappin, M. Effect of the Probiotic Enterococcus faecium SF68 on Presence of Diarrhea in Cats and Dogs Housed in an Animal Shelter. J. Vet. Intern. Med. 2011, 25, 856–860. [Google Scholar] [CrossRef]
  12. Strasser, B.; Geiger, D.; Schauer, M.; Gostner, J.M.; Gatterer, H.; Burtscher, M.; Fuchs, D. Probiotic Supplements Beneficially Affect Tryptophan-Kynurenine Metabolism and Reduce the Incidence of Upper Respiratory Tract Infections in Trained Athletes: A Randomized, Double-Blinded, Placebo-Controlled Trial. Nutrients 2016, 8, 752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Pollmann, M.; Nordhoff, M.; Pospischil, A.; Tedin, K.; Wieler, L.H. Effects of a probiotic strain of Enterococcus faecium on the rate of natural chlamydia infection in swine. Infect. Immun. 2005, 73, 4346–4353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gálvez, A.; Abriouel, H.; López, R.L.; Omar, N.B. Bacteriocin-based strategies for food biopreservation. Int. J. Food Microbiol. 2007, 120, 51–70. [Google Scholar] [CrossRef] [PubMed]
  15. Riley, M.A.; Gordon, D.M. The ecological role of bacteriocins in bacterial competition. In Trends in Microbiology; Elsevier Ltd.: Amsterdam, The Netherlands; Volume 7, pp. 129–133.
  16. José, M.; Burgos, G.; Pulido, R.P.; Aguayo, M.d.L.; Gálvez, A.; Lucas, R. The Cyclic Antibacterial Peptide Enterocin AS-48: Isolation, Mode of Action, and Possible Food Applications. Int. J. Mol. Sci. 2014, 15, 22706–22727. [Google Scholar]
  17. Shima, J.; Nisin, A. Nippon Shokuhin Kagaku Kogaku Kaishi. SJR 2008, 55, 37. [Google Scholar]
  18. Klaenhammer, T.R. Genetics of bacteriocins produced by lactic acid bacteria. FEMS Microbiol. Rev. 1993, 12, 39–85. [Google Scholar] [CrossRef]
  19. 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]
  20. Kumariya, R.; Garsa, A.K.; Rajput, Y.S.; Sood, S.K.; Akhtar, N.; Patel, S. Bacteriocins: Classification, synthesis, mechanism of action and resistance development in food spoilage causing bacteria. Microb. Pathog. 2019, 128, 171–177. [Google Scholar] [CrossRef]
  21. Cruz, V.L.; Ramos, J.; Melo, M.N.; Martinez-Salazar, J. Bacteriocin AS-48 binding to model membranes and pore formation as revealed by coarse-grained simulations. Biochim. Biophys. Acta Biomembr. 2013, 1828, 2524–2531. [Google Scholar] [CrossRef]
  22. Haas, W.; Gilmore, M.S. Molecular nature of a novel bacterial toxin: The cytolysin of Enterococcus faecalis. Med. Microbiol. Immunol. 1999, 187, 183–190. [Google Scholar] [CrossRef]
  23. Samyn, B.; Bueno, M.M.; Devreese, B.; Maqueda, M.; Galvez, A.; Valdivia, E.; Coyette, J.; van Beeumen, J. The cyclic structure of the enterococcal peptide antibiotic AS-48. FEBS Lett. 1994, 352, 87–90. [Google Scholar] [CrossRef] [Green Version]
  24. Stepper, J.; Shastri, S.; Loo, T.S.; Preston, J.C.; Novak, P.; Man, P.; Moore, C.H.; Havlicek, V.; Patchett, M.L.; Norris, G.E. Cysteine S -glycosylation, a new post-translational modification found in glycopeptide bacteriocins. FEBS Lett. 2011, 585, 645–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Eijsink, V.G.H.; Skeie, M.; Middelhoven, P.H.; Brurberg, M.B.; Nes, I.F. Comparative studies of class IIa bacteriocins of lactic acid bacteria. Appl. Environ. Microbiol. 1998, 64, 3275–3281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Aymerich, T.; Holo, H.; Håvarstein, L.S.; Hugas, M.; Garriga, M.; Nes, I.F. Biochemical and genetic characterization of enterocin A from Enterococcus faecium, a new antilisterial bacteriocin in the pediocin family of bacteriocins. Appl. Environ. Microbiol. 1996, 62, 1676–1682. [Google Scholar] [CrossRef] [Green Version]
  27. Hu, C.B.; Malaphan, W.; Zendo, T.; Nakayama, J.; Sonomoto, K. Enterocin X, a novel two-peptide bacteriocin from Enterococcus faecium KU-B5, has an antibacterial spectrum entirely different from those of its component peptides. Appl. Environ. Microbiol. 2010, 76, 4542–4545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Cintas, L.M.; Casaus, P.; Holo, H.; Hernandez, P.E.; Nes, I.F.; Håvarstein, L.S. Enterocins L50A and L50B, two novel bacteriocins from Enterococcus faecium L50, are related to staphylococcal hemolysins. J. Bacteriol. 1998, 180, 1988–1994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Casaus, P.; Nilsen, T.; Cintas, L.M.; Nes, I.F.; Hernández, P.E.; Holo, H. Enterocin B, a new bacteriocin from enterococcus faecium T136 which can act synergistically with enterocin A. Microbiology 1997, 143, 2287–2294. [Google Scholar] [CrossRef] [Green Version]
  30. Nilsen, T.; Nes, I.F.; Holo, H. Enterolysin A, a cell wall-degrading bacteriocin from Enterococcus faecalis LMG 2333. Appl. Environ. Microbiol. 2003, 69, 2975–2984. [Google Scholar] [CrossRef] [Green Version]
  31. Booth, M.C.; Bogie, C.P.; Sahl, H.G.; Siezen, R.J.; Hatter, K.L.; Gilmore, M.S. Structural analysis and proteolytic activation of Enterococcus faecalis cytolysin, a novel lantibiotic. Mol. Microbiol. 1996, 21, 1175–1184. [Google Scholar] [CrossRef]
  32. Maky, M.A.; Ishibashi, N.; Zendo, T.; Perez, R.H.; Doud, J.R.; Karmi, M.; Sonomoto, K. Enterocin F4-9, a novel O-linked glycosylated bacteriocin. Appl. Environ. Microbiol. 2015, 81, 4819–4826. [Google Scholar] [CrossRef] [Green Version]
  33. Cintas, L.M.; Casaus, P.; Håvarstein, L.S.; Hernández, P.E.; Nes, I.F. Biochemical and genetic characterization of enterocin P, a novel sec- dependent bacteriocin from Enterococcus faecium P13 with a broad antimicrobial spectrum. Appl. Environ. Microbiol. 1977, 63, 4321–4330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Criado, R.; Gutiérrez, J.; Martín, M.; Herranz, C.; Hernández, P.E.; Cintas, L.M. Immunochemical characterization of temperature-regulated production of enterocin L50 (EntL50A and EntL50B), enterocin P, and enterocin Q by Enterococcus faecium L50. Appl. Environ. Microbiol. 2006, 72, 7634–7643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. García, M.T.; Cañamero, M.M.; Lucas, R.; Omar, N.B.; Pulido, R.P.; Gálvez, A. Inhibition of Listeria monocytogenes by enterocin EJ97 produced by Enterococcus faecalis EJ97. Int. J. Food Microbiol. 2004, 90, 161–170. [Google Scholar] [CrossRef]
  36. Tiwari, S.K.; Srivastava, S. Purification and characterization of plantaricin LR14: A novel bacteriocin produced by Lactobacillus plantarum LR/14. Appl. Microbiol. Biotechnol. 2003, 79, 759–767. [Google Scholar] [CrossRef] [PubMed]
  37. Khan, H.; Flint, S.; Yu, P.; Yu, C.P.-L.; Group, B. Determination of the mode of action of enterolysin A, produced by Enterococcus faecalis B9510. J. Appl. Microbiol. 2013, 115, 484–494. [Google Scholar] [CrossRef] [PubMed]
  38. Sahl, H.G.; Bierbaum, G. Lantibiotics: Biosynthesis and biological activities of uniquely modified peptides from gram-positive bacteria. Annu. Rev. Microbiol. 1998, 52, 41–79. [Google Scholar] [CrossRef] [PubMed]
  39. Nes, I.F.; Diep, D.B.; Håvarstein, L.S.; Brurberg, M.B.; Eijsink, V.; Holo, H. Biosynthesis of bacteriocins in lactic acid bacteria. Antonie Van Leeuwenhoek Int. J. Gen. Mol. Microbiol. 1996, 70, 113–128. [Google Scholar] [CrossRef]
  40. Risøen, P.A.; Brurberg, M.B.; Eijsink, V.G.H.; Nes, I.F. Functional analysis of promoters involved in quorum sensing-based regulation of bacteriocin production in Lactobacillus. Mol. Microbiol. 2000, 37, 619–628. [Google Scholar] [CrossRef] [Green Version]
  41. Flynn, S.; van Sinderen, D.; Thornton, G.M.; Holo, H.; Nes, I.F.; Collins, J.K. Characterization of the genetic locus responsible for the production of ABP-118, a novel bacteriocin produced by the probiotic bacterium Lactobacillus salivarius subsp. salivarius UCC118. Microbiology 2002, 148, 973–984. [Google Scholar]
  42. Hoang, K.; Moore, B.S. Natural Metabolic Diversity Encoded by the Enterocin Biosynthesis Gene Cluster. J. Am. Chem. Soc. 2000, 122, 5415–5416. [Google Scholar]
  43. Rodali, V.P.; Lingala, V.K.; Karlapudi, A.P.; Indira, M.; Venkateswarulu, T.C.; Babu, D.J. Biosynthesis and potential applications of bacteriocins. J. Pure Appl. Microbiol. 2013, 7, 2933–2945. [Google Scholar]
  44. Furlaneto-Maia, L.; Ramalho, R.; Rocha, K.R.; Furlaneto, M.C. Antimicrobial activity of enterocins against Listeria sp. and other food spoilage bacteria. Biotechnol. Lett. 2020, 42, 797–806. [Google Scholar] [CrossRef] [PubMed]
  45. Le, T.N.; Do, T.H.; Nguyen, T.N.; Tran, N.T.; Enfors, S.O.; Truong, N.H. Expression and Simple Purification Strategy for the Generation of Antimicrobial Active Enterocin P from Enterococcus faecium Expressed in Escherichia coli ER2566. Iran. J. Biotechnol. 2014, 12, 17–25. [Google Scholar] [CrossRef] [Green Version]
  46. Nigutová, K.; Serenčová, L.; Piknová, M.; Javorský, P.; Pristaš, P. Heterologous expression of functionally active enterolysin A, class III bacteriocin from Enterococcus faecalis, in Escherichia coli. Protein Expr. Purif. 2008, 60, 20–24. [Google Scholar] [CrossRef] [PubMed]
  47. Daniela, M.C.C.; Barraza, E.; Colombo, N.S.R.; Galván, A.E.; Acuña, L.; Minahk, C.J.; Bellomio, A. New insights into enterocin CRL35: Mechanism of action and immunity revealed by heterologous expression in Escherichia coli. Mol. Microbiol. 2017, 105, 922–933. [Google Scholar]
  48. Klocke, M.; Mundt, K.; Idler, F.; Jung, S.; Backhausen, J.E. Heterologous expression of enterocin A, a bacteriocin from Enterococcus faecium, fused to a cellulose-binding domain in Escherichia coli results in a functional protein with inhibitory activity against Listeria. Appl. Microbiol. Biotechnol. 2005, 67, 532–538. [Google Scholar] [CrossRef] [PubMed]
  49. Arbulu, S.; Jiménez, J.J.; Gútiez, L.; Feito, J.; Cintas, L.M.; Herranz, C.; Hernández, P.E. Cloning and expression of synthetic genes encoding native, hybrid- and bacteriocin-derived chimeras from mature class IIa bacteriocins, by Pichia pastoris (syn. Komagataella spp.). Food Res. Int. 2019, 121, 888–899. [Google Scholar] [CrossRef]
  50. Hu, X.; Mao, R.; Zhang, Y.; Teng, D.; Wang, X.; Xi, D.; Wang, J. Biotechnical paving of recombinant enterocin A as the candidate of anti-Listeria agent. BMC Microbiol. 2014, 14, 220. [Google Scholar] [CrossRef] [Green Version]
  51. Jiménez, J.J.; Borrero, J.; Gútiez, L.; Arbulu, S.; Herranz, C.; Cintas, L.M.; Hernández, P.E. Use of Synthetic Genes for Cloning, Production and Functional Expression of the Bacteriocins Enterocin A and Bacteriocin E 50-52 by Pichia pastoris and Kluyveromyces lactis. Mol. Biotechnol. 2014, 56, 571–583. [Google Scholar] [CrossRef]
  52. Basanta, A.; Gómez-Sala, B.; Sánchez, J.; Diep, D.B.; Herranz, C.; Hernández, P.E.; Cintas, L.M. Use of the Yeast Pichia pastoris as an Expression Host for Secretion of Enterocin L50, a Leaderless Two-Peptide (L50A and L50B) Bacteriocin from Enterococcus faecium L50. Appl. Environ. Microbiol. 2010, 76, 3314–3324. [Google Scholar] [CrossRef] [Green Version]
  53. Gutiérrez, J.; Bourque, D.; Criado, R.; Choi, Y.J.; Cintas, L.M.; Hernández, P.E.; Míguez, C.B. Heterologous extracellular production of enterocin P from Enterococcus faecium P13 in the methylotrophic bacterium Methylobacterium extorquens. FEMS Microbiol. Lett. 2005, 248, 125–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Jiménez, J.J.; Diep, D.B.; Borrero, J.; Gútiez, L.; Arbulu, S.; Nes, I.F.; Herranz, C.; Cintas, L.M.; Hernández, P.E. Cloning strategies for heterologous expression of the bacteriocin enterocin A by Lactobacillus sakei Lb790, Lb. plantarum NC8 and Lb. casei CECT475. Microb. Cell Factories 2015, 14, 166. [Google Scholar] [CrossRef] [PubMed]
  55. Martín, M.; Gutiérrez, J.; Criado, R.; Herranz, C.; Cintas, L.M.; Hernández, P.E. Cloning, production and expression of the bacteriocin enterocin a produced by Enterococcus faecium PLBC21 in Lactococcus lactis. Appl. Microbiol. Biotechnol. 2007, 76, 667–675. [Google Scholar] [CrossRef] [PubMed]
  56. Perez, R.H.; Ishibashi, N.; Inoue, T.; Himeno, K.; Masuda, Y.; Sawa, N.; Zendo, T.; Wilaipun, P.; Leelawatcharamas, V.; Nakayama, J.; et al. Functional Analysis of Genes Involved in the Biosynthesis of Enterocin NKR-5-3B, a Novel Circular Bacteriocin. J. Bacteriol. 2016, 198, 291–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Line, J.E.; Svetoch, E.A.; Eruslanov, B.V.; Perelygin, V.V.; Mitsevich, E.V.; Mitsevich, I.P.; Levchuk, V.P.; Svetoch, O.E.; Seal, B.S.; Siragusa, G.R.; et al. Isolation and Purification of Enterocin E-760 with Broad Antimicrobial Activity against Gram-Positive and Gram-Negative Bacteria. Antimicrob. Agents Chemother. 2008, 52, 1094–1100. [Google Scholar] [CrossRef] [Green Version]
  58. Cotter, P.D.; Ross, R.P.; Hill, C. Bacteriocins—A viable alternative to antibiotics? Nat. Rev. Microbiol. 2013, 11, 95–105. [Google Scholar] [CrossRef]
  59. Cotter, P.D.; Hill, C.; Ross, R.P. Food microbiology: Bacteriocins: Developing innate immunity for food. Nat. Rev. Microbiol. 2005, 3, 777–788. [Google Scholar] [CrossRef]
  60. Cascales, E.; Buchanan, S.K.; Duché, D.; Kleanthous, C.; Lloubes, R.; Postle, K.; Riley, M.; Slatin, S.; Cavard, D. Colicin Biology. Microbiol. Mol. Biol. Rev. 2007, 71, 158–229. [Google Scholar] [CrossRef] [Green Version]
  61. Cotter, P.D. An ’Upp’-turn in bacteriocin receptor identification. Mol. Microbiol. 2014, 92, 1159–1163. [Google Scholar] [CrossRef]
  62. Herranz, C.; Chen, Y.; Chung, H.-J.; Cintas, L.M.; Hernández, P.E.; Montville, T.J.; Chikindas, M.L. Enterocin P Selectively Dissipates the Membrane Potential of Enterococcus faecium T136. Appl. Environ. Microbiol. 2001, 67, 1689–1692. [Google Scholar] [CrossRef] [Green Version]
  63. Cui, Y.; Luo, L.; Wang, X.; Lu, Y.; Yi, Y.; Shan, Y.; Liu, B.; Zhou, Y.; Lü, X. Mining, heterologous expression, purification, antibactericidal mechanism, and application of bacteriocins: A review. Compr. Rev. Food Sci. Food Saf. 2021, 20, 863–899. [Google Scholar] [CrossRef] [PubMed]
  64. Li, N.; Wang, Y.; You, C.; Ren, J.; Chen, W.; Zheng, H.; Liu, Z. Variation in Raw Milk Microbiota Throughout 12 Months and the Impact of Weather Conditions. Sci. Rep. 2018, 8, 2371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Khan, I.; Miskeen, S.; Khalil, A.T.; Phull, A.R.; Kim, S.J.; Oh, D.H. Foodborne pathogens: Staphylococcus aureus and listeria monocytogenes an unsolved problem of the food industry. Pak. J. Nutr. 2016, 15, 505–514. [Google Scholar] [CrossRef] [Green Version]
  66. Ananou, S.; Muñoz, A.; Bueno, M.M.; González-Tello, P.; Galvez, A.; Maqueda, M.; Valdivia, E. Evaluation of an enterocin AS-48 enriched bioactive powder obtained by spray drying. Food Microbiol. 2010, 27, 58–63. [Google Scholar] [CrossRef] [PubMed]
  67. Muñoz, A.; Ananou, S.; Galvez, A.; Bueno, M.M.; Rodriguez, A.; Maqueda, M.; Valdivia, E. Inhibition of Staphylococcus aureus in dairy products by enterocin AS-48 produced in situ and ex situ: Bactericidal synergism with heat. Int. Dairy J. 2007, 17, 760–769. [Google Scholar] [CrossRef]
  68. Lauková, A.; Mareková, M.; Javorský, P. Detection and antimicrobial spectrum of a bacteriocin-like substance produced by Enterococcus faecium CCM4231. Lett. Appl. Microbiol. 1993, 16, 257–260. [Google Scholar] [CrossRef]
  69. Lauková, A.; Czikková, S.; Dobránsky, T.; Burdová, O. Inhibition of Listeria monocytogenes and Staphylococcus aureus by enterocin CCM 4231 in milk products. Food Microbiol. 1999, 16, 93–99. [Google Scholar] [CrossRef]
  70. Lauková, A.; Vlaemynck, G.; Czikková, S. Effect of enterocin CCM 4231 on Listeria monocytogenes in Saint-Paulin cheese. Folia Microbiol. 2001, 46, 157–160. [Google Scholar] [CrossRef]
  71. Farías, M.E.; Farías, R.N.; Holgado, A.P.d.; Sesma, F. Purification and N-terminal amino acid sequence of enterocin CRL 35, a “pediocin-like” bacteriocin produced by Enterococcus faecium CRL 35. Lett. Appl. Microbiol. 1996, 22, 417–419. [Google Scholar] [CrossRef]
  72. Heredia, N.; García, S. Animals as sources of food-borne pathogens: A review. Anim. Nutr. 2018, 4, 250–255. [Google Scholar] [CrossRef]
  73. Hennekinne, J.-A.; Herbin, S.; Firmesse, O.; Auvray, F. European Food Poisoning Outbreaks Involving Meat and Meat-based Products. Procedia Food Sci. 2015, 5, 93–96. [Google Scholar] [CrossRef] [Green Version]
  74. Aymerich, T.; Garriga, M.; Ylla, J.; Vallier, J.; Monfort, J.M.; Hugas, M. Application of Enterocins as Biopreservatives against Listeria innocua in Meat Products. J. Food Prot. 2000, 63, 721–726. [Google Scholar] [CrossRef] [PubMed]
  75. Jofré, A.; Aymerich, T.; Monfort, J.M.; Garriga, M. Application of enterocins A and B, sakacin K and nisin to extend the safe shelf-life of pressurized ready-to-eat meat products. Eur. Food Res. Technol. 2008, 228, 159–162. [Google Scholar] [CrossRef]
  76. Ananou, S.; Garriga, M.; Hugas, M.; Maqueda, M.; Bueno, M.M.; Gálvez, A.; Valdivia, E. Control of Listeria monocytogenes in model sausages by enterocin AS-48. Int. J. Food Microbiol. 2005, 103, 179–190. [Google Scholar] [CrossRef] [PubMed]
  77. Ananou, S.; Baños, A.; Maqueda, M.; Martínez-Bueno, M.; Gálvez, A.; Valdivia, E. Effect of combined physico-chemical treatments based on enterocin AS-48 on the control of Listeria monocytogenes and Staphylococcus aureus in a model cooked ham. Food Control 2010, 21, 478–486. [Google Scholar] [CrossRef]
  78. Ananou, S.; Garriga, M.; Jofré, A.; Aymerich, T.; Galvez, A.; Maqueda, M.; Bueno, M.M.; Valdivia, E. Combined effect of enterocin AS-48 and high hydrostatic pressure to control food-borne pathogens inoculated in low acid fermented sausages. Meat Sci. 2010, 84, 594–600. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, G.; Wang, Y.; Gui, M.; Zheng, H.; Dai, R.; Li, P. Combined effect of high hydrostatic pressure and enterocin LM-2 on the refrigerated shelf life of ready-to-eat sliced vacuum-packed cooked ham. Food Control 2012, 24, 64–71. [Google Scholar] [CrossRef]
  80. Lauková, A.; Czikková, S.; Laczková, S.; Turek, P. Use of enterocin CCM 4231 to control Listeria monocytogenes in experimentally contaminated dry fermented Hornad salami. Int. J. Food Microbiol. 1999, 52, 115–119. [Google Scholar] [CrossRef]
  81. Al-Tayyar, N.A.; Youssef, A.M.; Al-Hindi, R.R. Edible coatings and antimicrobial nanoemulsions for enhancing shelf life and reducing foodborne pathogens of fruits and vegetables: A review. Sustain. Mater. Technol. 2020, 26, e00215. [Google Scholar] [CrossRef]
  82. Raybaudi-Massilia, R.M.; Mosqueda-Melgar, J.; Martín-Belloso, O. Edible alginate-based coating as carrier of antimicrobials to improve shelf-life and safety of fresh-cut melon. Int. J. Food Microbiol. 2008, 121, 313–327. [Google Scholar] [CrossRef]
  83. Lucas, R.; Grande, M.J.; Abriouel, H.; Maqueda, M.; Ben Omar, N.; Valdivia, E.; Cañamero, M.M.M.; Galvez, A. Application of the broad-spectrum bacteriocin enterocin AS-48 to inhibit Bacillus coagulans in canned fruit and vegetable foods. Food Chem. Toxicol. 2006, 44, 1774–1781. [Google Scholar] [CrossRef] [PubMed]
  84. Molinos, A.C.; Abriouel, H.; Omar, N.B.; Lucas, R.; Valdivia, E.; Gálvez, A. Inactivation of Listeria monocytogenes in raw fruits by enterocin AS-48. J. Food Prot. 2008, 71, 2460–2467. [Google Scholar] [CrossRef] [PubMed]
  85. Molinos, A.C.; Abriouel, H.; López, R.L.; Omar, N.B.; Valdivia, E.; Gálvez, A. Inhibition of Bacillus cereus and Bacillus weihenstephanensis in raw vegetables by application of washing solutions containing enterocin AS-48 alone and in combination with other antimicrobials. Food Microbiol. 2008, 25, 762–770. [Google Scholar]
  86. Molinos, A.C.; Abriouel, H.; Ben Omar, N.; Valdivia, E.; López, R.L.; Maqueda, M.; Cañamero, M.M.; Gálvez, A. Effect of Immersion Solutions Containing Enterocin AS-48 on Listeria monocytogenes in Vegetable Foods. Appl. Environ. Microbiol. 2005, 71, 7781–7787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Grande, M.J.; Lucas, R.; Valdivia, E.; Abriouel, H.; Maqueda, M.; Ben Omar, N.; Cañamero, M.M.M.; Galvez, A. Stability of enterocin AS-48 in fruit and vegetable juices. J. Food Prot. 2005, 68, 2085–2094. [Google Scholar] [CrossRef]
  88. Grande, M.J.; Lucas, R.; Abriouel, H.; Ben Omar, N.; Maqueda, M.; Martínez-Bueno, M.; Martínez-Cañamero, M.; Valdivia, E.; Gálvez, A. Control of Alicyclobacillus acidoterrestris in fruit juices by enterocin AS-48. Int. J. Food Microbiol. 2005, 104, 289–297. [Google Scholar] [CrossRef]
  89. Grande, M.J.; Abriouel, H.; López, R.L.; Valdivia, E.; Ben Omar, N.; Cañamero, M.M.M.; Gálvez, A. Efficacy of enterocin AS-48 against bacilli in ready-to-eat vegetable soups and purees. J. Food Prot. 2007, 70, 2339–2345. [Google Scholar] [CrossRef]
  90. Pulido, R.P.; Toledo, J.; Grande, M.J.; Gálvez, A.; Lucas, R. Analysis of the effect of high hydrostatic pressure treatment and enterocin AS-48 addition on the bacterial communities of cherimoya pulp. Int. J. Food Microbiol. 2015, 196, 62–69. [Google Scholar] [CrossRef]
  91. Anacarso, I.; de Niederhäusern, S.; Iseppi, R.; Sabia, C.; Bondi, M.; Messi, P. Anti-listerial activity of chitosan and Enterocin 416K1 in artificially contaminated RTE products. Food Control 2011, 22, 2076–2080. [Google Scholar] [CrossRef]
  92. Viedma, P.M.; Abriouel, H.; Omar, N.B.; López, R.L.; Gálvez, A. Effect of enterocin EJ97 against Geobacillus stearothermophilus vegetative cells and endospores in canned foods and beverages. Eur. Food Res. Technol. 2009, 230, 513–519. [Google Scholar] [CrossRef]
  93. Issouffou, C.; Suwansri, S.; Salaipeth, L.; Domig, K.J.; Hwanhlem, N. Synergistic effect of essential oils and enterocin KT2W2G on the growth of spoilage microorganisms isolated from spoiled banana peel. Food Control 2018, 89, 260–269. [Google Scholar] [CrossRef]
  94. Lihui, D.; Xingrong, J.; Jian, Y.; Rong, H.; Lifeng, W.; Wei, S.; Qin, L. Enterococcus Durable and Its Bacteriocin That Can Inhibit Listeria. China Patent CN103305445B, 21 January 2015. [Google Scholar]
  95. Xingrong, J.; Lihui, D.; Xinquan, C.; Xueyou, W.; Jian, Y.; Rong, H.; Ying, H.X.; Lingping, L. A Durable Enterococcus Mutant, Encoding Gene and Its Application. China Patent CN105037509A, 11 November 2015. [Google Scholar]
  96. Guicheng, H.; Xueqin, W.; Lijie, Y. A High-Yielding Bacteriocin Enterococcus faecalis and Its Application. China Patent CN104593294A, 6 May 2015. [Google Scholar]
  97. Junrui, W.; Rina, W.; Xiaorui, W.; Xiqing, Y.; Dongbing, T. A Bacteriocin-Producing Enterococcus faecium and Its Application. China Patent CN105779346B, 17 May 2019. [Google Scholar]
  98. Chen, H.; Hoover, D.G. Bacteriocins and Their Food Applications. Compr. Rev. Food Sci. Food Saf. 2003, 2, 82–100. [Google Scholar]
Figure 1. (Adapted from [43]) Schematic Diagram (A): The biosynthesis pathway of several class I enterocins (lanthipeptides): (1) formation of preenterocin; (2) modification of preenterocin by LanB and LanC and translocation of preenterocin by LanT; (3) histidine protein kinase (HPK) senses the presence of enterocin and autophosphorylates; (4) the phosphoryl group (P) is transferred to the response regulator (RR); (5) the transcription of regulated genes is activated by RR; (6) mature enterocin is processed and released under the regulation of LanP, LanI and Lan EFG. Schematic Diagram (B): The biosynthesis pathway of class II enterocins: (1) formation of the preenterocin and prepeptide of the induction factor (IF); (2) the precursors are processed and translocated by the ABC transporter and then mature enterocin and IF are released; (3) HPK senses the presence of autophosphorylates and IF; (4) the phosphoryl group (P) is transferred to the response regulator (RR); (5) the transcription of regulated genes is activated by RR; (6) producer immunity is achieved.
Figure 1. (Adapted from [43]) Schematic Diagram (A): The biosynthesis pathway of several class I enterocins (lanthipeptides): (1) formation of preenterocin; (2) modification of preenterocin by LanB and LanC and translocation of preenterocin by LanT; (3) histidine protein kinase (HPK) senses the presence of enterocin and autophosphorylates; (4) the phosphoryl group (P) is transferred to the response regulator (RR); (5) the transcription of regulated genes is activated by RR; (6) mature enterocin is processed and released under the regulation of LanP, LanI and Lan EFG. Schematic Diagram (B): The biosynthesis pathway of class II enterocins: (1) formation of the preenterocin and prepeptide of the induction factor (IF); (2) the precursors are processed and translocated by the ABC transporter and then mature enterocin and IF are released; (3) HPK senses the presence of autophosphorylates and IF; (4) the phosphoryl group (P) is transferred to the response regulator (RR); (5) the transcription of regulated genes is activated by RR; (6) producer immunity is achieved.
Molecules 27 02258 g001
Figure 2. The antibacterial mechanism of enterocins (adapted from [59]). Enterocins can be divided into different groups according to their mode of action: (a) Several class I enterocins have two antimicrobial methods. They can bind to lipid II and prevent peptidoglycan synthesis. Furthermore, they can also use lipid II to initiate membrane insertion and form pores. Class II enterocins generally have an amphiphilic helical structure, allowing them to be inserted into the cell membrane and causing the death of target cell. Class III enterocins, such as Enterolysin A, can degrade the sensitive cells by hydrolyzing the specific peptide bonds of cell wall. (b) A number of enterocins control the target bacteria by interfering with gene expression.
Figure 2. The antibacterial mechanism of enterocins (adapted from [59]). Enterocins can be divided into different groups according to their mode of action: (a) Several class I enterocins have two antimicrobial methods. They can bind to lipid II and prevent peptidoglycan synthesis. Furthermore, they can also use lipid II to initiate membrane insertion and form pores. Class II enterocins generally have an amphiphilic helical structure, allowing them to be inserted into the cell membrane and causing the death of target cell. Class III enterocins, such as Enterolysin A, can degrade the sensitive cells by hydrolyzing the specific peptide bonds of cell wall. (b) A number of enterocins control the target bacteria by interfering with gene expression.
Molecules 27 02258 g002
Table 1. Classification of Enterocins.
Table 1. Classification of Enterocins.
ClassificationCharactersExamples
Class II aposttranslationally modified; lanthipeptidesCytolysin [31]
I bposttranslationally modified; head-to-tail cyclized peptides;Enterocin AS-48 [16]
I cposttranslationally modified; sactibiotics
I dposttranslationally modified; linear azol(in)e-containing peptides
I eposttranslationally modified; glycocinsEnterocin F4-9 [32]
I fposttranslationally modified; lasso peptides
Class IIII aunmodified bacteriocins; pediocin-like bacteriocinsEnterocin A [25];
Enterocin P [33]
II bunmodified bacteriocins; two-peptide bacteriocinsEnterocin X [27]
II cunmodified bacteriocins; leaderless bacteriocinsEnterocin L50 [28];
Enterocin Q [34];
Enterocin EJ97 [35];
Enterocin RJ-11 [36]
II dunmodified bacteriocins; non-pediocin-like, single-peptide bacteriocinsEnterocin B [29]
Class III large molecular weight; heat labileEnterolysin A [37]
Table 2. Heterologous expression of enterocins in different systems.
Table 2. Heterologous expression of enterocins in different systems.
Expression SystemsEnterocins Producing StrainExpression HostPlasmidReferences
E. coliEnterocin PE. faecium P13E. coli ER2566pTWIN1[45]
Enterolysin AE. faecalis II/1E. coli SG13009pQE-30 UA[46]
Enterocin CRL35E. mundtii CRL35BL21(DE3)pACYCDuet-1[47]
Enterocin A and BE. faecium ATB197aBL21(DE3)pET37b(+)[48]
Yeast Enterocin HF and enterocin CRL35E. faeciumP. pastoris X-33pPICZαA[49]
Enterocin AE. faecium CTC492; E. faecium T136P. pastoris X-33; K. lactis GG799EApPICZαA; pKLAC2[50,51]
L50A and L50BE. faecium L50P. pastoris X-33pPICZαA[52]
Enterocin PE. faecium P13P. pastoris X-33pPICZαA[53]
LactobacillusEnterocin AE. faecium T136;
E. faecium PLBC21
L. casei CECT475;
L. casei IL1403;
L. casei NZ9000
pSIP411UAI;
pMG36c
[54,55]
Ent53BE. faecium NKR-5-3L. casei NZ9000pNK-B[56]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, Y.; Pang, X.; Wu, Y.; Liu, X.; Zhang, X. Enterocins: Classification, Synthesis, Antibacterial Mechanisms and Food Applications. Molecules 2022, 27, 2258. https://doi.org/10.3390/molecules27072258

AMA Style

Wu Y, Pang X, Wu Y, Liu X, Zhang X. Enterocins: Classification, Synthesis, Antibacterial Mechanisms and Food Applications. Molecules. 2022; 27(7):2258. https://doi.org/10.3390/molecules27072258

Chicago/Turabian Style

Wu, Yajing, Xinxin Pang, Yansha Wu, Xiayu Liu, and Xinglin Zhang. 2022. "Enterocins: Classification, Synthesis, Antibacterial Mechanisms and Food Applications" Molecules 27, no. 7: 2258. https://doi.org/10.3390/molecules27072258

Article Metrics

Back to TopTop