Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Bioengineering of a Lactococcus lactis subsp. lactis strain enhances nisin production and bioactivity

  • Roxana Portieles ,

    Contributed equally to this work with: Roxana Portieles, Hongli Xu

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Hongli Xu ,

    Contributed equally to this work with: Roxana Portieles, Hongli Xu

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Feng Chen,

    Roles Investigation, Validation

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Jingyao Gao,

    Roles Formal analysis, Investigation

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Lihua Du,

    Roles Investigation, Methodology, Visualization

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Xiangyou Gao,

    Roles Project administration, Supervision, Validation, Visualization

    Affiliation Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China

  • Carlos Borroto Nordelo,

    Roles Investigation, Methodology, Project administration, Resources, Supervision, Visualization

    Affiliation ‎VBS Biotec SA, Yuc., México

  • Qiulin Yue,

    Roles Data curation, Formal analysis, Investigation

    Affiliation State Key Laboratory of Biobased Material and Green Papermaking, Shandong Provincial Key Lab of Microbial Engineering, Qilu University of Technology (Shandong Academic of Science), Jinan, People’s Republic of China

  • Lin Zhao,

    Roles Data curation, Formal analysis, Investigation, Methodology, Software

    Affiliation State Key Laboratory of Biobased Material and Green Papermaking, Shandong Provincial Key Lab of Microbial Engineering, Qilu University of Technology (Shandong Academic of Science), Jinan, People’s Republic of China

  • Nayanci Portal Gonzalez,

    Roles Investigation, Methodology, Validation

    Affiliation School of Biological Science and Technology, University of Jinan, Jinan, Shandong, People’s Republic of China

  • Ramon Santos Bermudez,

    Roles Data curation, Formal analysis, Investigation, Software, Validation, Visualization

    Affiliation School of Biological Science and Technology, University of Jinan, Jinan, Shandong, People’s Republic of China

  • Orlando Borrás-Hidalgo

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    orlando@yotabio.com

    Affiliations Joint R and D Center of Biotechnology, RETDA, YOTABIO-ENGINEERING CO., LTD., Rizhao, Shandong, P.R. China, State Key Laboratory of Biobased Material and Green Papermaking, Shandong Provincial Key Lab of Microbial Engineering, Qilu University of Technology (Shandong Academic of Science), Jinan, People’s Republic of China

Abstract

Lactococcus lactis subsp. lactis is a food bacterium that has been utilized for decades in food fermentation and the development of high-value industrial goods. Among these, nisin, which is produced by several strains of L. lactis subsp. lactis, plays a crucial role as a food bio-preservative. The gene expression for nisin synthesis was evaluated using qPCR analysis. Additionally, a series of re-transformations of the strain introducing multiple copies of the nisA and nisRK genes related to nisin production were developed. The simultaneous expression of nisA and nisZ genes was used to potentiate the effective inhibition of foodborne pathogens. Furthermore, qPCR analysis indicated that the nisA and nisRK genes were expressed at low levels in wild-type L. lactis subsp. lactis. After several re-transformations of the strain with the nisA and nisRK genes, a high expression of these genes was obtained, contributing to improved nisin production. Also, co-expression of the nisA and nisZ genes resulted in extremely effective antibacterial action. Hence, this study would provide an approach to enhancing nisin production during industrial processes and antimicrobial activity.

Introduction

The control of foodborne pathogens constitutes an important issue in the food industry. Diseases produced by these kinds of pathogens are disseminated rapidly [1]. Although diverse actions were deployed to decrease the infections [2], foodborne diseases remain prevalent worldwide [3]. To limit the potential risks, many methods of managing foodborne pathogens have been developed [2]. Bacteriocins can be used as effective alternatives for food preservatives [4, 5].

Lactococcus lactis is used to produce different fermented milk products. L. lactis was previously classified as a Streptococcus species. In 1985, it was reclassified as a member of the Lactococcus genus [6] and grouped into three subspecies: L. lactis, L. lactis subsp. hordniae, and L. lactis subsp. cremoris [7]. This Gram-positive bacterium is spherical, non-sporulating, homolactate, and facultative [8, 9]. L. lactis was used in food fermentation for centuries. Also, L. lactis produces acid, which helps preserve food. Some strains improve on this preservation feature by producing bacteriocins.

The antimicrobial peptide nisin is composed of 34 amino acids that are produced during the growth of various L. lactis strains [10]. Generally, nisin inhibits and prevents the spread of Gram-positive bacteria [1113]. This constitutes an important element to be used as an excellent food preservative and additive in the food industry [1418].

Other likely applications of nisin are related to the biomedical field [1922]. Numerous studies have highlighted the use of nisin in therapeutic approaches. For example, nisin had immunomodulatory effects similar to those human defense peptides [23]. Besides, nisin has been shown to be effective in the treatment of carcinoma [21, 24]. Recently, a solution of nisin Z supplemented with EDTA could be used as antibiotic treatment in diabetic foot infections [22]. An interesting study provided important evidence that nisin Z could be also safely used topically as disinfectant at recommended concentrations without skin toxicity [20].

The sophisticated biosynthesis of nisin has been associated with a group of eleven genes. These genes are grouped into structural genes (nisA), transport (nisT), post-translational modifications (nisB, nisC), extracellular precursor processing (nisP), immunity (nisI, nisFEG), and the regulatory genes (nisR, nisK) [25, 26]. Enhanced nisin yield appears to require a better understanding of the nisin biosynthetic process and its metabolic regulation. The rate and level of nisin production slightly improved when immunity genes were expressed to increase nisin resistance [27].

The analysis of nisin biosynthesis shows the way for carrying out genetic modifications designed to raise nisin yield. The genetic modifications were centralized on the issues that limit the production of nisin. Innovative genetic approaches have increased nisin production with L. lactis strains expressing the aox1, pfk13, and pkaC genes [28]. These recombinant nisin producers induced oxidative respiration and glycolytic activity. Additionally, nisin Z production was improved through the use of protoplast fusion and genome shuffle in L. lactis ssp. lactis YF11 [29]. By simultaneously expressing the hdeAB, Idh, and murF genes, the nisin production of the producer L. lactis was improved, as was its acidic tolerance [30]. Herewith, over-expression of the asnH gene increased the acidic tolerance in the nisin producer L. lactis [31].

Previous research suggests that an increase in nisin production was facilitated by a boosting in the copy number of the crucial genes involved in nisin biosynthesis [3236]. Additionally, enhancing the copy number of the genes nisRK and nisFEG, which are responsible for regulating and resisting nisin biosynthesis, significantly increased nisin production [36]. Most of the precedent works were done through a simple genetic transformation and selection of the most promising strains with high production of nisin. However, it is unknown about the impact of the genetic re-transformation on L. lactis subsp. lactis strains several times with genes involved in nisin synthesis. Previously, the genetic re-transformation was associated with the significant increase of proteins in bacteria and yeast. The re-transformation of Clostridium acetobutylicum bacterium with genes expressing stress proteins resulted in a more robust phenotype and improved n-butanol tolerance during fermentations [37]. While the yeast Pichia pastoris significantly increased its ability to synthesize insulin when the same approach was used [38].

The use of bacteriocins could be hindered by restricted activity, high concentrations of foodborne pathogens, and antimicrobial resistance. When the bacteriocins are combined or used in conjunction with other antimicrobial agents, multiple studies have demonstrated additive effects [15, 17, 18]. Thus, this strategy is a good alternative for effective pathogen inhibition. The combination of nisin, carvacrol, and citric acid inhibited Gram-positive and Gram-negative foodborne pathogens [39]. The activity of nisin was enhanced with the addition of EDTA against L. monocytogenes and Escherichia coli [40]. The combination of polymyxin, colistin, and nisin increased the activity against the biofilm formation of Pseudomonas aeruginosa [41]. Also, a L. lactis strain that co-expressed nisin and leucocin C demonstrated potent antibacterial activity. This strain had a broad activity range and a good bacteriostatic capacity against Gram-positive foodborne pathogens [42]. Curiously, the biological activity of a L. lactis strain co-expressing simultaneously the nisA and nisZ genes was not previously evaluated.

Therefore, the aim of this study was the characterization and evaluation of serial genetic re-transformations (RT) to enhance the nisin production in the L. lactis subsp. lactis. Additionally, we constructed a L. lactis subsp. lactis strain co-expressing the nisA and nisZ genes with enhanced antimicrobial activity. This work describes how it is possible to achieve a significant increase in nisin production through serial genetic re-transformations in the L. lactis subsp. lactis strain with high activity.

Materials and methods

Bacterial strains, plasmids, and culture conditions

The Escherichia coli DH5α (Invitrogen, USA) was grown in an LB (Luria-Bertani) medium (Difco) at 37°C for cloning and the construction of plasmids. The Lactococcus lactis subsp. lactis CICC 6242 strain used in the experiments was supplied by the China Center of Industrial Culture Collection (Beijing, People’s Republic of China). The pGEM-T Easy Vector (Promega, Madison, USA) and pMG36e vector (Addgene, MA, USA) were used in the cloning and recombinant expression experiments. The different fermentations of L. lactis subsp. lactis were done in 250-mL flasks containing 50 mL of M17 medium (Difco) containing 0.5% of glucose. The L. lactis subsp. lactis was incubated by shaking at 100 rpm at 30°C for 18 hours without control of pH. The erythromycin (Sangon, Shanghai, China) was used at 250 μg/mL for E. coli DH5α and 5 μg/mL for L. lactis subsp. lactis strains, respectively. The cell growth (O.D. 600), pH and nisin titers (IU/ml) parameters were evaluated every two hours until 18 hours of fermentation.

Preparation of recombinant strains

Chromosomal DNA from L. lactis subsp. lactis CICC 6242 strain producing nisin A was extracted using a “Wizard Genomic DNA Purification” kit (Promega, Madison, USA). The nisA and nisRK genes were amplified through a polymerase chain reaction using the primers listed in Table 1 [36]. The PCRs were performed using the following reagents: 2.5 μL of DNA polymerase buffer, 2.5 μL of dNTPs, 0.5 μL of MgCl2, 0.5 μL of each primer, 0.1 μL of DNA polymerase (Promega, Madison, USA) at 50 ng/μL and enough water in a final volume of 12.5 μL. Amplifications were performed in a Bio-Rad thermocycler T100TM (Bio Rad, USA) with the following program: an initial denaturation step at 94 °C for 5 min, followed by 35 denaturation cycles at 94 °C for 1 min, annealing (nisA: 62 °C and nisRK 60 °C), extension at 72 °C for 1 min and a final extension step at 72 °C for 7 min. The amplified products were stained with a mix of 2X Blue/Orange Loading Dye (Promega, Madison, USA), loading buffer (1X) and 1X SYBr green Nucleic Acid Gel Stain (Sangon Biotech, Shanghai) dye and were separated by electrophoresis on 1% agarose gels. The sequence integrity from nisA and nisRK genes were evaluated by sequencing. Herewith, the nisZ gene was synthesized (Shanghai RealGene Bio-Tech, Inc).

The genes were cloned one by one into in pGEM-T Easy Vector (Promega, Madison, USA) and finally in pMG36e vector (Addgene, MA, USA) under p32 promoter to the recombinant expression using PCR Cloning Kit (Qiagen, Germany). The recombinant plasmids were named as pMG36e:nisA (nisA), pMG36e:nisZ (nisZ) and pMG36e:nisRK (nisRK). The recombinant plasmids were first introduced in E. coli DH5 cells and cultured on LB agar plates with 250 μg/mL of erythromycin. Restriction enzyme digestions and sequencing analyses were used to examine the DNA plasmids. To obtain the electro-competent L. lactis cells, the bacterium was cultivated overnight at 30 °C until an O.D.600 of 0.7. The cells were resuspended in 1/100 volume of electroporation solution (0.5 M sucrose and 10% glycerol) and kept on ice, one the cells were washed twice with ice-cold washing solution (0.5 M sucrose and 10% glycerol). The confirmed recombinant plasmids were electro-transformed in the L. lactis subsp. lactis strain using a Gene Pulser device (Bio Rad, USA) set at 2.2 kV, 200, and 25°F. Likewise, the transformants pMG36e:nisA, pMG36e:nisZ, and pMG36e:nisRK were selected on M17 medium containing 5 μg/mL erythromycin. The plasmids from the transformant strains were further analyzed by DNA sequencing.

The strategies of re-transformation and expression of nisA and nisRK genes in the same strain were as follow: a) genetic re-transformation was done in the original L. lactis subsp. lactis strain with the nisA and nisRK genes and the colonies were selected by erythromycin resistance and the number of copies of the genes introduced regarding the number of copies of these genes in the original strain; b) independent clones of L. lactis subsp. lactis resistant to erythromycin and with the highest copies numbers of nisA and nisRK genes were selected, respectively; c) two more subsequent transformations were carried out using in each step the clones with the highest copies numbers as selection criteria (pMG36e:nisA RT and pMG36e:nisRK RT), since the antibiotic resistance of the plasmid is the same (S1 Fig).

In order to get both nisins expressed in the same strain, the original and multiple copies of the nisA L. lactis subsp. lactis strains were also transformed with the nisZ gene. The selection of the clones was based on antibiotic resistance, the number of copies of the nisA and nisZ genes, relative expression, RNA sequencing, and LC-MS/MS analysis (S2 Fig). To test their biological activity, clones expressing nisin A, Z, or both in the same strain were selected. In addition, the genetic stability of expression plasmid vector pMG36e in L. lactis subsp. lactis was evaluated by culturing transformants pMG36e:nisA, pMG36e:nisA RT, pMG36e:nisRK and pMG36e:nisRK RT in 100 μl of GM17 broth without erythromycin selection. The L. lactis subsp. lactis was maintained at an exponential phase for 120 generations. The L. lactis subsp. lactis was grown until upper log phase, and 100 μl of culture was inoculated into fresh 100 ml GM17 broth. This method was repeated until the cells had achieved roughly 120 generations. The L. lactis subsp. lactis was then plated on GM17 agar and incubated at 30°C for 18 hours. A total of 100 colonies were selected at random and subcultured on GM17 agar with erythromycin. Finally, the percentage of colonies that retained the plasmid was calculated.

Analysis of gene copy number of transformed strains

Total DNA from all the transformed and wild-type L. lactis subsp. lactis strains were extracted using a “Wizard Genomic DNA Purification” kit (Promega, Madison, USA). Real-time PCR was done using a Rotor-Gene Q machine (Qiagen, Hilden, Germany) with the QuantiTect SYBR Green PCR Kit (Qiagen, Germany). For absolute quantification of the nisA and nisRK genes in transformed and wild-type L. lactis subsp. lactis strains, two standard curves at different dilutions (1:10; 1:100; 1:1000, and 1:10000) were established using the genes previously cloned in pGEM-T Easy Vector (Promega, Madison, USA). The data were determined in terms of copies of the nisA and nisRK genes / µL of qPCR reaction. The primers used in the experiments are listed in Table 1 [36]. Real-time PCR conditions were as follows: an initial 95°C denaturation step for 15 min followed by denaturation for 15 sec at 95°C, annealing for 30 sec at 55°C, and extension for 30 sec at 72°C for 40 cycles. Data were analyzed by one-way analysis of variance using GraphPad Prism 5.0 (GraphPad Software, Inc, California). Significant differences among means were determined by Tukey’s Multiple Comparison Test least significant difference mean separation at P < 0.05.

Transcriptional analysis by quantitative real-time PCR

The original L. lactis subsp. lactis and transformed strains were incubated in shaking at 100 rpm at 30°C during the fermentation. The samples for RNA isolation were taken at 0; 12 and 18 hours of fermentation. Total RNAs were isolated with Trizol (Invitrogen, Carlsbad, CA, USA) according to the manufacturer’s instructions. The cDNA was synthesized using the SuperScript III reverse transcriptase kit (Invitrogen, Carlsbad, CA, USA). The primers used in the experiments are listed in Table 1) [36]. For each condition, reactions were done in triplicate in three separate experiments. To standardize cycle threshold (Ct) results, the 16S rRNA gene was used as an internal control. Quantitative real-time PCR was done using a Rotor-Gene Q machine (Qiagen, Hilden, Germany) with the QuantiTect SYBR Green PCR Kit (Qiagen, Germany). Real-time PCR conditions were as follows: an initial 95°C denaturation step for 15 min followed by denaturation for 15 sec at 95°C, annealing for 30 sec at 55°C, and extension for 30 sec at 72°C for 40 cycles. The relative levels of genes were expressed as ‘mean normalized expression’ data using Q-Gene software [43]. The Q-Gene software application is a tool for dealing with complex quantitative real-time PCR experiments on a large scale, significantly accelerating and rationalizing experimental setup, data analysis, and data management while ensuring maximum reproducibility. Data were analyzed by one-way analysis of variance using GraphPad Prism 5.0 (GraphPad Software, Inc, California). Significant differences among means were determined by Tukey’s Multiple Comparison Test least significant difference mean separation at P < 0.05.

Nisin quantification and bioactivity assay

The nutrient broth (NB) medium (peptones 10 g/L, beef extract 1 g/L, yeast extract 2 g/L, sodium chloride 5 g/L, and pH 6.8) for Micrococcus luteus was used during the bioassay agar plates [44]. Before boiling and sterilizing, 0.75% Bacto agar (Difco) and 1% of Tween 20 were added to the medium composition. After autoclaving, the medium was inoculated with 1% of the 24-h culture of M. luteus. The concentration of the microorganism was approximately 108 colony-forming units/mL. The bioassay agar was placed aseptically into sterile Petri dishes (100 X 15 mm) and allowed to solidify for 3 hours. Four holes were done on each plate using a 7-mm outer diameter stainless steel borer. A well was filled with an aliquot (100 µL) of standard nisin solution and samples, and the bioassay agar plates were incubated at 30 °C for 24 hours. The diameter of the inhibitory zone was measured in the nisin-sensitive microorganism. Three replication were used in the bioassay. Multiple comparisons using Tukey’s test were done.

Also, the nisin quantification was estimated by agar diffusion activity assays and expressed in international units / mL (IU/mL) [44]. A stock nisin solution (1,000 IU/mL) was prepared by adding 0.025 g of commercial nisin106 IU/g (Sigma Aldrich) into 25 mL of the sterile diluent solution of 0.02 N HCl: L. lactis fermentation medium (9:1 by volume), which consisted of (per L): 80 g of glucose, 10 g of peptone, 10 g of yeast extract, 10 g of KH2PO4, 2 g of NaCl, and 0.2 g of MgSO4.7H2O. Standard nisin solutions of 500, 400, 300, 200, 100, 50, 25, 10, 5, and 0 IU/mL were prepared using the 1,000 IU/mL nisin stock solution to construct the standard curve. To establish a standard curve, the diameters of the inhibition zones were plotted against the log10 concentrations of nisin. A line regression equation was determined for a standard curve. The diameter of the inhibition zone obtained from the nisin standard solution was correlated with the bioassay of sensitivity [44].

Chloroform extraction and LC-MS/MS analysis

The transformed L. lactis subsp. lactis and wild-type strains overnight culture were inoculated in M17 medium (Difco) and incubated in shaking at 100 rpm at 30°C for 18 hours. Bacterium cells were pelleted at 7500 g for 15 min in a refrigerated centrifuge (12 °C) and the supernatant was collected. Chloroform was added to the supernatant (1:1), stirred vigorously for 20 min, and centrifuged at 10400 g (12 °C) for 30 min. The precipitated and interface solids were collected. The solids were dried in a chemical hood overnight. Finally, the solids were suspended in 1 ml of Tris buffer (0.1 mol/L, 5±10 ml, pH 7.0) with agitation overnight at 8 °C. A membrane filter (0.20 m) was used to filter the extraction, which was suspended in 1 ml of Tris buffer. The filtered sample solution was then used for LC-MS/MS analysis. An HPLC system (1200 series, Agilent, Santa Clara, CA) and triple quadruple mass spectrometer (Agilent 6410), were used to identify and quantify the analyses. A CW-C18 column (50x2 mm, 3 m; Imtakt, Portland, OR) was used for the chromatographic analysis, and the column oven temperature was maintained at 40 °C. The determination of nisin A and nisin Z content in the fermentation was done quantitatively using LC-MS/MS analysis with positive mode electrospray ionization and the multiple reaction-monitoring parameters. Data were analyzed with Agilent 6410 Quantitative analysis version (B.02.01) analyst data-processing software.

Results

Characterization of L. lactis subsp. lactis wild-type strain

The fermentation profiles of the L. lactis subsp. lactis wild-type strain were determined at different time points. The nisin titer was 324 IU/ml after 18 hours of culture. Further, the cell growth reached a maximum of 1.81 optical density after 12 hours. Additionally, there was a strong reduction in pH which had a value of 4.5 upon finishing the fermentation process (Fig 1).

thumbnail
Fig 1. Evaluation of different parameters during the fermentation of the L. lactis subsp. lactis strain wild type.

(A) Temporal growth profile. (B) pH of the culture. (C) Nisin A production. The bars represent means of three independent replicates ± standard error (P<0.05).

https://doi.org/10.1371/journal.pone.0281175.g001

Also, the expression of the genes related to nisin production in the wild-type strain of L. lactis subsp. lactis was evaluated. The majority of the nisin biosynthetic genes had variable transcription levels. The nisI, nisF, and nisG genes showed higher levels of expression. However, the nisR and nisK genes had low expression in the different time points analyzed. Also, the nisA gene showed reduced expression (Fig 2). According to these results, the genes involved in nisin pre-peptide synthesis (nisA) and two-component signal transduction (nisRK) were selected for further genetic transformation in the L. lactis subsp. lactis wild-type strain.

thumbnail
Fig 2. Transcriptional analysis of genes involved in the biosynthesis of nisin in the L. lactis subsp. lactis strain using qRT-PCR.

For each condition, reactions were done in triplicate in three separate experiments. To standardize cycle threshold (Ct) results, the 16S rRNA gene was used as an internal control. The relative expression was expressed as ‘mean normalized expression’ data using Q-Gene software [43]. The bars represent means of three independent replicates ± standard error (P<0.05).

https://doi.org/10.1371/journal.pone.0281175.g002

Serial genetic re-transformations to nisin production

We designed an approach based on a re-transformation of the nisA and nisRK genes into the wild-type L. lactis subsp. lactis strain. The transcription levels of nisA and nisRK genes were higher in modified L. lactis subsp. lactis strains generated following a single transformation than in the original strain (Fig 3). However, the relative expression was significantly higher in re-transformed L. lactis subsp. lactis strains compared with the single transformed and wild-type strains, respectively (Fig 3).

thumbnail
Fig 3. Transcriptional analysis of the nisA, nisR, and nisK genes in the wild-type and engineered L. lactis subsp. lactis strains using qRT-PCR.

The relative expression values were determined regarding the 16S rRNA gene used as reference internal gene. The relative levels of genes were expressed as ‘mean normalized expression’ data using Q-Gene software [43]. The bars represent means of three independent replicates ± standard error (P<0.05).

https://doi.org/10.1371/journal.pone.0281175.g003

The growth of the engineered L. lactis subsp. lactis strains was determined through their optical densities at 600 nm. The engineered L. lactis subsp. lactis pMG36e:nisA, pMG36e:nisA RT, pMG36e:nisRK, and pMG36e:nisRK RT strains displayed a similar cell growth density compared with the wild-type strain. There was no effect of the over-expression of these genes on cell growth (Fig 4A). A high and significant number of copies of the nisA genes were obtained in the engineered pMG36e:nisA RT and pMG36e:nisRK RT strains compared with the single transformed and wild-type strains, respectively (Fig 4B).

thumbnail
Fig 4. Characterization of the L. lactis subsp. lactis wild-type and engineered strains.

(A) cell growth was evaluated through of optical density OD600 after 18 hours of fermentation. (B) the value data were determined in terms of copies of the nisA and nisRK genes / µL of qPCR reaction. (C) the nisin A quantification was determined by activity assay in the L. lactis subsp. lactis wild-type and engineered strains. The bars represent means of three independent replicates ± standard error (P<0.05).

https://doi.org/10.1371/journal.pone.0281175.g004

The nisin productivity of the original strain and the engineered strains (pMG36e:nisA, pMG36e:nisA RT, pMG36e:nisRK, and pMG36e:nisRK RT) were evaluated in shake flasks after 18 h using agar diffusion activity assays. The wild-type and transformed strains after single and three-fold transformations with empty plasmid pMG36e exhibited similar nisin production, suggesting this had no effect on the nisin production. Though, the pMG36e:nisA and pMG36e:nisRK strains showed a high production of nisin, with 442.5 IU/ml and 369.2 IU/ml, respectively, after 18 hours, compared with the wild-type strain (169.3 IU/ml) (Fig 4C).

Among these engineered strains, pMG36e:nisA RT (1111.5 IU/ml) and pMG36e:nisRK RT (1041.5 IU/ml) had the greatest increments in nisin production (Fig 4C). These results stated that the strains genetically modified through re-transformation had a positive effect on nisin yield. Also, with or without erythromycin selection pressure, the pMG36e genetically stable percentage of retention in L. lactis subsp. lactis was about 98 percent after 120 generations of continuous incubation (S3 Fig).

Co-expressing nisA and nisZ genes

To evaluate the advantages of the co-expression of nisA and nisZ genes in terms of biological activity, we compared them with the original strain. Antimicrobial assays were done with engineered L. lactis subsp. lactis pMG36e:nisA, pMG36e:nisZ, and pMG36e:nisA + pMG36e:nisZ constructs. The L. lactis subsp. lactis strain producing nisin A and nisin Z was verified by RNA sequencing and mass spectrometry, respectively. For biological activity, a strain selected with similar transcript expression levels from the nisA and nisZ genes was used. Additionally, LC-MS/MS analysis showed the production of nisin A and Z in the same selected strain at similar concentrations. All the strains inhibited the M. luteus indicator strain through antagonistic bioactivity assays. However, the strain expressing the nisA and nisZ genes had the highest inhibitory effect on the indicator strain (Fig 5). In terms of bacteriostatic activity against M. luteus, co-expression of nisin A and Z was superior to the original strain.

thumbnail
Fig 5. Antimicrobial activity of L. lactis subsp. lactis strains expressing nisA, nisZ and nisA + nisZ genes against M. luteus.

The plates’ holes were filled with 150 ml of pasteurized supernatant from overnight cultures of L. lactis subsp. lactis strains. (A) Inhibition area; (B) inhibition diameter in millimeters. The bars represent means of three independent replicates ± standard error (P<0.05).

https://doi.org/10.1371/journal.pone.0281175.g005

Discussion

The parameters of cell growth and pH during fermentation remained adequate in the wild-type L. lactis subsp. lactis strain. A reduction in pH typically associated with lactic acid production was observed at the end of fermentation, which is closely related to nisin production. Although the wild-strain of L. lactis subsp. lactis produced certain levels of nisin at the end of fermentation by shake flask culture, the levels were low compared to previous work where other wild-strain of L. lactis subsp. lactis were used [36, 45, 46]. To look for the causes of these low levels, a molecular characterization through qPCR was developed with the aim of determining how the set of genes involved in nisin production were expressed.

Different approaches for understanding nisin biosynthesis and its metabolism have been evaluated [25, 4648]. It is important to note that each of these genes should not be analyzed in a separate context. However, the nisR and nisK gene expression was significantly low in the wild-type L. lactis subsp. lactis strain. The nisA and nisRK genes display a significant role in the synthesis and regulation of nisin [36, 45, 49]. It was evident that the expression of these genes was limited, which could explain in some way the low production of nisin obtained. This has been demonstrated in other strains of L. lactis subsp. lactis how the increasing of these genes could enhance nisin production [46]. Specifically, when the produced nisin induces the two-component signal transduction, the nisRK genes display an important role [25, 26, 47]. Although, we focused on these genes. Others genes involved in the nisin biosynthetic could be evaluated in the future.

Given this, increasing the expression of these genes could, in principle, improve the production of nisin in the wild-type strain. Consequently, we increased the expression of these genes by introducing multiple copies of the nisA and nisRK genes in L. lactis subsp. lactis. To increase the expression of nisin, the nisA and nisRK genes were cloned in a high-copy-number vector. The resulting plasmids, pMG36e:nisA and pMG36e:nisRK, were introduced into L. lactis subsp. lactis through single- and three-fold genetic re-transformation. The results showed how it is possible to first increase gene expression and then increase the production of nisin. The highest expression profiles were detected in the engineered strains obtained through the re-transformation of the wild-type strain compared with the single transformation wild-type strain. Our data support the idea that high expression levels of the nisA and nisRK genes in the engineered strains increase nisin production. We successfully achieved a high production of nisin without any influence over the growth and stability of the engineered L. lactis subsp. lactis strains in the shake flasks. Interesting, in L. lactis subsp. lactis, pMG36e demonstrated high genetic stability. The genetic stability of pMG36e was not affected by erythromycin selection pressure. Previously, the stability of the pMG36e plasmid was demonstrated in the L. lactis M4 strain [50].

On the other hand, the cell growth of high-yield engineered strains showed no differences compared with the control strains. There was no toxicity effect produced by the presence of multiple copies of these genes, including the engineered strains obtained by re-transformation. The results provide an approach for improving nisin production. Additional studies related to the optimization of fermentation conditions in engineered strains in scale production should be developed.

Diverse efforts have been developed to increase nisin activity [42, 51, 52]. Combinations with other bacteriocins have allowed an increase in activity [42, 52]. Additionally, the combination with compounds that weaken the cell wall of Gram-negative bacteria such as E. coli has increased the spectrum of action [42, 53, 54]. In this context, the differences between nisin A and nisin Z are minimal [14]. Within these differences, the different amino acids in position 27 and the ability of nisin Z to achieve a higher level of inhibition with respect to nisin A, mainly due to its ability to have a greater degree of diffusion in agar with respect to nisin A, are the most important [14].

An interesting point was how L. lactis subsp. lactis would be expressing nisin A and nisin Z simultaneously. The results suggest that there is a potentiation of activity when the two nisins are produced in the same strain of L. lactis subsp. lactis, compared to those that are expressed separately. There was a positive effect between them, which contributes to better activity in terms of inhibition activity.

Several analyses have shown that nisin has additive or synergistic effects when used with other antimicrobial compounds [15, 17, 18, 40, 41]. However, there was no evidence available about the antibacterial activity of the L. lactis subsp. lactis strain co-expressing nisin A and nisin Z. Our findings show that the co-expression of these two bacteriocins with minor variations resulted in extremely effective antibacterial action against M. luteus.

Finally, many efforts have been made to seek strategies that allow an increase in the production of nisin. Within these, the increase in the expression of genes related to the biosynthesis of nisin was extensively addressed. Our work, although focused on this issue, introduces the strategy of genetic retransformation for an increase in the number of significant copies, without affecting stability and with a marked increase in the yield of nisin. On the other hand, it was interesting to evaluate the biological effect of the expression of the nisA and nisZ genes in the same strain of L. lactis. It is well known for the amino acid difference at position 27 and the greater agar diffusion capacity of nisin Z compared to nisin A. This was demonstrated; however, it was curious to observe how this strain, which produced the same concentration of nisin A and Z, had a greater inhibitory effect. Further studies will be necessary to understand the reason for this increase in activity and behavior during scale-up in fermenters. Additionally, optimization in fermenter conditions would be important to introduce these L. lactis strains genetically modified in a production flow.

Supporting information

S1 Fig. Schematic representation of the strategies of re-transformation and expression of nisA and nisRK genes in L. lactis.

https://doi.org/10.1371/journal.pone.0281175.s001

(TIF)

S2 Fig. Schematic representation of the strategies of co-expression of nisA and nisZ genes in L. lactis.

https://doi.org/10.1371/journal.pone.0281175.s002

(TIF)

S3 Fig. Evaluation of genetic stability of the pMG36e plasmid expressing of nisA and nisRK genes in L. lactis.

https://doi.org/10.1371/journal.pone.0281175.s003

(TIF)

References

  1. 1. Käferstein FK, Motarjemi Y, Bettcher DW. Foodborne disease control: a transnational challenge. Emerg Infect Dis. 1997; 3(4):503–10. pmid:9368787
  2. 2. Gandhi M, Chikindas ML. Listeria: A foodborne pathogen that knows how to survive. Int J Food Microbiol. 2007; 113(1):1–15. pmid:17010463
  3. 3. Velusamy V, Arshak K, Korostynska O, Oliwa K, Adley C. An overview of foodborne pathogen detection: in the perspective of biosensors. Biotechnol Adv. 2010; 28(2):232–54. pmid:20006978
  4. 4. Roberts RF, Zottola EA. Shelf-life of pasteurized process cheese spreads made from cheddar cheese manufactured with a nisin-producing starter culture. J Dairy Sci. 1993; 76(7):1829–36. pmid:8345121
  5. 5. Lucera A, Costa C, Conte A, Del Nobile MA. Food applications of natural antimicrobial compounds. Front Microbiol. 2012; 3:287. pmid:23060862
  6. 6. Song AA, In LLA, Lim SHE, Rahim RA. A review on Lactococcus lactis: from food to factory. Microb Cell Fact. 2017; 16(1):55. Erratum in: Microb Cell Fact. 2017; 16(1):139.
  7. 7. Parapouli M, Delbès-Paus C, Kakouri A, Koukkou AI, Montel MC, Samelis J. Characterization of a wild, novel nisin a-producing Lactococcus strain with an L. lactis subsp. cremoris genotype and an L. lactis subsp. lactis phenotype, isolated from Greek raw milk. Appl Environ Microbiol. 2013; 79(11):3476–84.
  8. 8. Garrigues C, Loubiere P, Lindley ND, Cocaign-Bousquet M. Control of the shift from homolactic acid to mixed-acid fermentation in Lactococcus lactis: predominant role of the NADH/NAD+ ratio. J Bacteriol. 1997; 179(17):5282–7.
  9. 9. Duwat P, Sourice S, Cesselin B, Lamberet G, Vido K, Gaudu P, et al. Respiration capacity of the fermenting bacterium Lactococcus lactis and its positive effects on growth and survival. J Bacteriol. 2001; 183(15):4509–16.
  10. 10. Liu W, Hansen JN. Some chemical and physical properties of nisin, a small-protein antibiotic produced by Lactococcus lactis. Appl Environ Microbiol. 1990; 56(8):2551–8.
  11. 11. Severina E, Severin A, Tomasz A. Antibacterial efficacy of nisin against multidrug-resistant Gram-positive pathogens. J Antimicrob Chemother. 1998; 41(3):341–7. pmid:9578160
  12. 12. Periago PM, Moezelaar R. Combined effect of nisin and carvacrol at different pH and temperature levels on the viability of different strains of Bacillus cereus. Int J Food Microbiol. 2001; 68(1–2):141–8.
  13. 13. Hampikyan H. Efficacy of nisin against Staphylococcus aureus in experimentally contaminated sucuk, a Turkish-type fermented sausage. J Food Prot. 2009; 72(8):1739–43.
  14. 14. Williams GC, Delves-Broughton J. in Encyclopedia of Food Sciences and Nutrition (Second Edition) 2003
  15. 15. Dai Z, Han L, Li Z, Gu M, Xiao Z, Lu F. Combination of Chitosan, Tea Polyphenols, and Nisin on the Bacterial Inhibition and Quality Maintenance of Plant-Based Meat. Foods. 2022; 11(10):1524. pmid:35627094
  16. 16. Maresca D, Mauriello G. Development of Antimicrobial Cellulose Nanofiber-Based Films Activated with Nisin for Food Packaging Applications. Foods. 2022; 11(19):3051. pmid:36230127
  17. 17. Wu M, Ma Y, Dou X, Zohaib Aslam M, Liu Y, Xia X, et al. A review of potential antibacterial activities of nisin against Listeria monocytogenes: the combined use of nisin shows more advantages than single use. Food Res Int. 2023; 164:112363.
  18. 18. Chen H, Ji PC, Qi YH, Chen SJ, Wang CY, Yang YJ, et al. Inactivation of Pseudomonas aeruginosa biofilms by thymoquinone in combination with nisin. Front Microbiol. 2023; 13:1029412.
  19. 19. Shin JM, Gwak JW, Kamarajan P, Fenno JC, Rickard AH, Kapila YL. Biomedical applications of nisin. J Appl Microbiol. 2016; 120(6):1449–65. pmid:26678028
  20. 20. Soltani S, Boutin Y, Couture F, Biron E, Subirade M, Fliss I. In vitro assessment of skin sensitization, irritability and toxicity of bacteriocins and reuterin for possible topical applications. Sci Rep. 2022; 12(1):4570.
  21. 21. Khazaei Monfared Y, Mahmoudian M, Cecone C, Caldera F, Zakeri-Milani P, Matencio A, et al. Stabilization and Anticancer Enhancing Activity of the Peptide Nisin by Cyclodextrin-Based Nanosponges against Colon and Breast Cancer Cells. Polymers (Basel). 2022; 14(3):594. pmid:35160583
  22. 22. Serrano I, Alhinho B, Cunha E, Tavares L, Trindade A, Oliveira M. Bacteriostatic and Antibiofilm Efficacy of a Nisin Z Solution against Co-Cultures of Staphylococcus aureus and Pseudomonas aeruginosa from Diabetic Foot Infections. Life (Basel). 2023; 13(2):504.
  23. 23. Kindrachuk J, Jenssen H, Elliott M, Nijnik A, Magrangeas-Janot L, Pasupuleti M, et al. Manipulation of innate immunity by a bacterial secreted peptide: lantibiotic nisin Z is selectively immunomodulatory. Innate Immun. 2013; 19(3):315–27. pmid:23109507
  24. 24. Joo NE, Ritchie K, Kamarajan P, Miao D, Kapila YL. Nisin, an apoptogenic bacteriocin and food preservative, attenuates HNSCC tumorigenesis via CHAC1. Cancer Med. 2012; 1(3):295–305. pmid:23342279
  25. 25. Kuipers OP, Beerthuyzen MM, de Ruyter PG, Luesink EJ, de Vos WM. Autoregulation of nisin biosynthesis in Lactococcus lactis by signal transduction. J Biol Chem. 1995; 270(45):27299–304.
  26. 26. Ra SR, Qiao M, Immonen T, Pujana I, Saris PEJ. Genes responsible for nisin synthesis, regulation and immunity form a regulon of two operons and are induced by nisin in Lactoccocus lactis N8. Microbiology. 1996; 142 (Pt 5):1281–1288.
  27. 27. Kim WS, Hall RJ, Dunn NW. Improving nisin production by increasing nisin immunity/resistance genes in the producer organism Lactococcus lactis. Appl. Microbiol Biotechnol. 1998, 50, 429–433
  28. 28. Papagianni M, Avramidis N. Engineering the central pathways in Lactococcus lactis: functional expression of the phosphofructokinase (pfk) and alternative oxidase (aox1) genes from Aspergillus niger in Lactococcus lactis facilitates improved carbon conversion rates under oxidizing conditions. Enzyme Microb Technol. 2012; 51(3):125–30.
  29. 29. Zhang YF, Liu SY, Du YH, Feng WJ, Liu JH, Qiao JJ. Genome shuffling of Lactococcus lactis subspecies lactis YF11 for improving nisin Z production and comparative analysis. J Dairy Sci. 2014; 97(5):2528–41.
  30. 30. Zhang J, Caiyin Q, Feng W, Zhao X, Qiao B, Zhao G, et al. Enhance nisin yield via improving acid-tolerant capability of Lactococcus lactis F44. Sci Rep. 2016; 6:27973.
  31. 31. Hao P, Liang D, Cao L, Qiao B, Wu H, Caiyin Q, et al. Promoting acid resistance and nisin yield of Lactococcus lactis F44 by genetically increasing D-Asp amidation level inside cell wall. Appl Microbiol Biotechnol. 2017; 101(15):6137–6153.
  32. 32. Kim WS, Hall RJ, Dunn NW. The effect of nisin concentration and nutrient depletion on nisin production of Lactococcus lactis. Appl Microbiol Biotechnol. 1997; 48(4):449–53.
  33. 33. Cheigh CI, Choi HJ, Park H, Kim SB, Kook MC, Kim TS, et al. Influence of growth conditions on the production of a nisin-like bacteriocin by Lactococcus lactis subsp. lactis A164 isolated from kimchi. J Biotechnol. 2002; 95(3):225–35.
  34. 34. Simşek O, Akkoç N, Con AH, Ozçelik F, Saris PE, Akçelik M. Continuous nisin production with bioengineered Lactococcus lactis strains. J Ind Microbiol Biotechnol. 2009; 36(6):863–71.
  35. 35. Simşek O, Con AH, Akkoç N, Saris PE, Akçelik M. Influence of growth conditions on the nisin production of bioengineered Lactococcus lactis strains. J Ind Microbiol Biotechnol. 2009; 36(4):481–90.
  36. 36. Ni ZJ, Zhang XY, Liu F, Wang M, Hao RH, Ling PX, et al. Effect of Co-overexpression of Nisin Key Genes on Nisin Production Improvement in Lactococcus lactis LS01. Probiotics Antimicrob Proteins. 2017; 9(2):204–212.
  37. 37. Mann MS, Dragovic Z, Schirrmacher G, Lütke-Eversloh T. Over-expression of stress protein-encoding genes helps Clostridium acetobutylicum to rapidly adapt to butanol stress. Biotechnol Lett. 2012; 34(9):1643–9.
  38. 38. Mansur M, Cabello C, Hernández L, País J, Varas L, Valdés J, et al. Multiple gene copy number enhances insulin precursor secretion in the yeast Pichia pastoris. Biotechnol Lett. 2005; 27(5):339–45.
  39. 39. Campion A, Morrissey R, Field D, Cotter PD, Hill C, Ross RP. Use of enhanced nisin derivatives in combination with food-grade oils or citric acid to control Cronobacter sakazakii and Escherichia coli O157:H7. Food Microbiol. 2017; 65:254–263.
  40. 40. Branen JK, Davidson PM. Enhancement of nisin, lysozyme, and monolaurin antimicrobial activities by ethylene diamine tetra acetic acid and lactoferrin. Int. J. Food Microbiol. 2004, 90, 63–74. pmid:14672831
  41. 41. Field D, Seisling N, Cotter PD, Ross RP, Hill C. Synergistic Nisin-Polymyxin Combinations for the Control of Pseudomonas Biofilm Formation. Front Microbiol. 2016 Oct 26;7:1713.
  42. 42. Fu Y, Mu D, Qiao W, Zhu D, Wang X, Liu F, et al. Co-expression of Nisin Z and Leucocin C as a Basis for Effective Protection Against Listeria monocytogenes in Pasteurized Milk. Front Microbiol. 2018 Mar 23;9:547.
  43. 43. Muller PY, Janovjak H, Miserez AR, Dobbie Z. Processing of gene expression data generated by quantitative real-time RT-PCR. Biotechniques. 2002 Jun;32(6):1372–4, 1376, 1378–9. Erratum in: Biotechniques. 2002 Sep;33(3):514. pmid:12074169
  44. 44. Pongtharangkul T, Demirci A. Evaluation of agar diffusion bioassay for nisin quantification. Appl Microbiol Biotechnol. 2004 Aug;65(3):268–72. pmid:14963617
  45. 45. Cheigh CI, Park H, Choi HJ, Pyun YR. Enhanced nisin production by increasing genes involved in nisin Z biosynthesis in Lactococcus lactis subsp. lactis A164. Biotechnol Lett. 2005 Feb;27(3):155–60.
  46. 46. Khelissa S, Chihib NE, Gharsallaoui A. Conditions of nisin production by Lactococcus lactis subsp. lactis and its main uses as a food preservative. Arch Microbiol. 2021 Mar;203(2):465–480.
  47. 47. de Ruyter PG, Kuipers OP, Beerthuyzen MM, van Alen-Boerrigter I, de Vos WM. Functional analysis of promoters in the nisin gene cluster of Lactococcus lactis. J Bacteriol. 1996 Jun;178(12):3434–9.
  48. 48. Kleerebezem M, Quadri LE, Kuipers OP, de Vos WM. Quorum sensing by peptide pheromones and two-component signal-transduction systems in Gram-positive bacteria. Mol Microbiol. 1997 Jun;24(5):895–904. pmid:9219998
  49. 49. Kong W, Lu T. Cloning and optimization of a nisin biosynthesis pathway for bacteriocin harvest. ACS Synth Biol. 2014 Jul 18;3(7):439–45. pmid:24847677
  50. 50. Noreen N, Hooi WY, Baradaran A, Rosfarizan M, Sieo CC, Rosli MI, et al. Lactococcus lactis M4, a potential host for the expression of heterologous proteins. Microb Cell Fact. 2011 Apr 26;10:28.
  51. 51. Yoneyama F, Fukao M, Zendo T, Nakayama J, Sonomoto K. Biosynthetic characterization and biochemical features of the third natural nisin variant, nisin Q, produced by Lactococcus lactis 61–14. J Appl Microbiol. 2008 Dec;105(6):1982–90.
  52. 52. Piper C, Hill C, Cotter PD, Ross RP. Bioengineering of a Nisin A-producing Lactococcus lactis to create isogenic strains producing the natural variants Nisin F, Q and Z. Microb Biotechnol. 2011 May;4(3):375–82.
  53. 53. Abarca RL, Medina J, Alvarado N, Ortiz PA, Carrillo López B. Biodegradable gelatin-based films with nisin and EDTA that inhibit Escherichia coli. PLoS One. 2022 Mar 10;17(3):e0264851.
  54. 54. Field D, Baghou I, Rea MC, Gardiner GE, Ross RP, Hill C. Nisin in Combination with Cinnamaldehyde and EDTA to Control Growth of Escherichia coli Strains of Swine Origin. Antibiotics (Basel). 2017 Dec 12;6(4):35.