Next Article in Journal
Risk Factors Associated with Toxoplasma gondii in Patients with Cardiovascular Diseases from Western Romania
Next Article in Special Issue
Transmission and Persistence of Infant Gut-Associated Bifidobacteria
Previous Article in Journal
Distinct Gut Microbial Enterotypes and Functional Dynamics in Wild Striped Field Mice (Apodemus agrarius) across Diverse Populations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of Probiotic Properties and Whole-Genome Analysis of Lactobacillus johnsonii N5 and N7 Isolated from Swine

1
Jiangsu Co-Innovation Center for Prevention and Control of Important Animal Infectious Diseases and Zoonoses, Yangzhou University, Yangzhou 225009, China
2
Jiangsu Key Laboratory of Zoonosis, Yangzhou University, Yangzhou 225009, China
3
Joint International Research Laboratory of Agriculture and Agri-product Safety of the Ministry of Education, Yangzhou University, Yangzhou 225009, China
4
Key Laboratory of Prevention and Control of Biological Hazard Factors (Animal Origin) for Agrifood Safety and Quality, Ministry of Agriculture of China, Yangzhou University, Yangzhou 225009, China
5
College of Animal Science and Technology, Yangzhou University, Yangzhou 225009, China
*
Authors to whom correspondence should be addressed.
Microorganisms 2024, 12(4), 672; https://doi.org/10.3390/microorganisms12040672
Submission received: 21 February 2024 / Revised: 22 March 2024 / Accepted: 25 March 2024 / Published: 28 March 2024
(This article belongs to the Special Issue Beneficial Microbes and Gastrointestinal Microbiota 2.0)

Abstract

:
In our previous microbiome profiling analysis, Lactobacillus (L.) johnsonii was suggested to contribute to resistance against chronic heat stress-induced diarrhea in weaned piglets. Forty-nine L. johnsonii strains were isolated from these heat stress-resistant piglets, and their probiotic properties were assessed. Strains N5 and N7 exhibited a high survival rate in acidic and bile environments, along with an antagonistic effect against Salmonella. To identify genes potentially involved in these observed probiotic properties, the complete genome sequences of N5 and N7 were determined using a combination of Illumina and nanopore sequencing. The genomes of strains N5 and N7 were found to be highly conserved, with two N5-specific and four N7-specific genes identified. Multiple genes involved in gastrointestinal environment adaptation and probiotic properties, including acidic and bile stress tolerance, anti-inflammation, CAZymes, and utilization and biosynthesis of carbohydrate compounds, were identified in both genomes. Comparative genome analysis of the two genomes and 17 available complete L. johnsonii genomes revealed 101 genes specifically harbored by strains N5 and N7, several of which were implicated in potential probiotic properties. Overall, this study provides novel insights into the genetic basis of niche adaptation and probiotic properties, as well as the genome diversity of L. johnsonii.

1. Introduction

Lactic acid bacteria (LAB) are generally defined as a group of Gram-positive, non-spore-forming bacteria that are capable of producing lactic acid and exhibit negative catalase activity [1]. Numerous members of the LAB exhibit probiotic characteristics, with a notable emphasis on those affiliated with the genus Lactobacillus (taxonomy lineage Firmicutes, Bacilli, Lactobacillales, Lactobacillaceae) constituting a core group in this context. Probiotic members of the Lactobacillus genus can colonize the intestine and maintain the balance of intestinal microbiota, which includes inhibiting the growth of harmful bacteria in the intestine through various mechanisms such as producing organic acids, bacteriocin, and other antibacterial substances [2]. Furthermore, they exhibit a close association with the cells located in the lower regions of the intestinal mucosa and then benefit the host by participating in immune cell regulation and activation [3].
Among the species affiliated with the genus Lactobacillus, L. johnsonii is mainly distributed in the gastrointestinal tract of both humans and non-human animals [4]. L. johnsonii has been isolated from the intestines of piglets [5], dogs [6], chickens [7], calves [8], mice [9], bees [10], and humans [11]. It usually exhibits resistance against the corrosive effects of gastric acid and bile compounds, enabling it to traverse the gastrointestinal tract and adhere effectively [12]. L. johnsonii is regarded as a safe probiotic microorganism, holding a Generally Recognized as Safe (GRAS) status, and is included in the Qualified Presumption of Safety (QPS) list [13]. Many members of L. johnsonii exhibit several probiotic characteristics, including the maintenance of metabolic stability in the gastrointestinal tract, inhibition of in vivo growth and proliferation of pathogenic bacteria, and enhancement of animal growth performance. Given these probiotic characteristics, several L. johnsonii probiotic strains have been extensively utilized as probiotics in feed additives and probiotic preparations [14,15].
With the recent advancements in high-throughput sequencing technologies, notably long-read sequencing technologies, it has become feasible to acquire complete genome sequences of bacterial strains [16]. This capability enhances the precise and confident identification of genes associated with probiotic properties and other traits through genome mining, as well as genes linked to evolution, including those related to host adaptation [17,18]. L. johnsonii was suggested to be positively associated with the ability to resist chronic heat stress-induced diarrhea in weaned piglets, as revealed by comparative microbiome profiling analysis in our previous study [19]. Then, a total of 49 L. johnsonii strains were isolated from the feces of these heat-stress-tolerant piglets. A representative L. johnsonii strain, N5, was selected for in vivo experiments, and this strain was demonstrated to exhibit protective effects against dextran sulfate sodium (DSS)-induced colitis of the host by strengthening intestinal barrier function and balancing the response of intestinal innate and adaptive immunity [19]. In the current study, we conducted in vitro assessments of the antagonistic activity against pathogenic bacteria and the tolerance to gastric acid and bile of the L. johnsonii strains. Strain N5, along with another strain, N7, demonstrated notable antagonistic activity against Salmonella and exhibited high tolerance to acid and bile stresses. Complete genome sequences of strains N5 and N7 were determined by integrative analysis of Illumina short-read and nanopore long-read sequencing data. The genomic contents that potentially contributed to the intestinal colonization and functionality of L. johnsonii were identified. We further performed comparative genomic analyses of strains N5 and N7 with available complete genomes of L. johnsonii to discover shared and distinctive characteristics of N5 and N7, as well as to gain a comprehensive understanding of genome diversity and evolution of L. johnsonii.

2. Materials and Methods

2.1. Isolation of Lactic Acid Bacteria from Swine Feces

The LAB were isolated from the feces of weaned piglets that exhibited resistance to chronic heat stress (i.e., no diarrhea symptom) [19] through a dilution and plate coating method. Specifically, 2 g of each swine fecal sample was weighed and then added to a sterilized tube that contained 5 mL of sterilized PBS. The mixture was thoroughly mixed using vortex oscillation. The homogenized fecal samples were diluted to 10−4, 10−5, and 10−6, and 100 μL of each dilution were evenly spread on MRS plates, and incubated upside down at 37 °C for 24 h. Colonies were evenly distributed without any apparent adhesion phenomenon. A total of 30–300 colonies with different morphologies and colors were picked for subsequent cross-purification. Each strain was purified three times, followed by Gram staining and microscopic examination to ensure complete purification of the strain. In case of inconsistent morphology observed under microscopic examination, repeated line purification was performed until consistent morphology was achieved. The LAB was incubated in the corresponding medium at 37 °C for 18 to 24 h and subsequently sub-cultured under the same conditions for all subsequent assays.

2.2. Taxonomic Identification of LAB

The genomic DNA of each isolate was extracted using a bacterial genomic DNA extraction kit (Tiangen Biotech Co., Ltd. Beijing, China). Subsequently, the bacterial 16S rDNA sequence was amplified with the forward primer 27F (5′-AGAGTTTGATCCTGGCTCAG-3′) and reverse primer 1492R (5′-GGYTACCTTGTTACGACTT-3′). The PCR products were further verified through 1% agarose gel electrophoresis and subjected to Sanger sequencing (Genscript, Nanjing, China). The taxonomic affiliation of these strains was accomplished by comparing their sequences with 16S rDNA reference sequences deposited in the NCBI rDNA database using the BLASTn program. Out of the 57 obtained LAB strains, 49 strains were identified to be affiliated with L. johnsonii and subjected to further experiments.

2.3. Antagonistic Activity against Salmonella

The antibacterial activity was assessed following the methodology outlined by Leite et al. [20] with certain modifications. Ten microliters of each overnight L. johnsonii strain was spotted onto MRS agar and incubated for 24 h at 37 °C. Subsequently, the overnight culture of strain Salmonella (S.) Typhimurium SL1344 was inoculated into an LB semi-solid medium containing 0.8% agar (precooled to ambient temperature) at a ratio of 1:100 (v/v). The resulting mixture was evenly poured onto the MRS agar plates containing LAB strains. After solidification, the plates were incubated overnight at 37 °C. The presence of clear halo zones surrounding the L. johnsonii colony (measured in millimeters) was indicative of the antagonistic activity of the L. johnsonii strain. The strains that showed no evident halo information or a halo measuring less than 1 mm were recorded as “−”, and those that showed the presence of a distinct growth inhibition zone surrounding spots larger than 1 mm were scored as “+” (positive). Those with an inhibition zone ranging from 2 to 5 mm around the colony were recorded as “++”.

2.4. Bile and Acidic Tolerance

The L. johnsonii strains were cultured overnight in MRS broth at 37 °C and then inoculated into fresh MRS broth with a pH of 2.5 or 0.3% bovine bile salt at a 5% inoculation rate. After uniform distribution using a pipette, the strains were incubated at 37 °C for 4 h. A 0.1 mL aliquot of bacterial suspension was collected before culturing (0 h) and at two different time points (2 h and 4 h) for subsequent gradient dilution. Following appropriate dilutions (10−3, 10−4), a 0.1 mL diluent was applied onto MRS agar plates for enumeration. The tolerance rate of L. johnsonii strains to acidic and bile stresses (%) was determined by applying the following formula: (the number of viable bacteria at 2 and 4 h (CFU/mL)/the number of viable bacteria at 0 h (CFU/mL)) × 100%.

2.5. Analytical Profile Index (API) Experiment

Single colonies of L. johnsonii N5 and N7 freshly cultured on MRS solid medium were selected and used to prepare a bacterial suspension with a turbidity of 2 McFarland in API 50 CHL medium (BioMérieux, Lyon, France). Subsequently, 5 mL of distilled water was poured into the honeycomb wells of the culture tray to create a humid environment. API 50 test strips were placed in the tray, and then bacterial suspension was added to each well of the biochemical test strip. The wells were then sealed with liquid paraffin and the tray was incubated at 37 °C for 24 h in a static incubator. Based on the instruction table, the biochemical reaction results were interpreted for N5 and N7.

2.6. Whole Genome Sequencing and Genome Assembly

The whole-genome sequencing was conducted on an Illumina HiSeq4000 platform (short reads) and a nanopore platform (long reads). High-quality long reads (with a Q score greater than 7 and length exceeding 1 kb) generated by the nanopore sequencing were assembled using Flye (v2.8.3) [21]. The generated assemblies were subsequently refined using NextPolish (v2) [22] with the Illumina-generated short reads as inputs. The genomic sequences of L. johnsonii N5 and N7 have been deposited at GenBank with the accession numbers CP136013 and CP136012.

2.7. Comparative Genomics Analysis

Seventeen completely sequenced L. johnsonii genome sequences available in the NCBI database (as of 10 March 2022) (Table 1) were downloaded, and the genomes of N5 and N7 and 17 downloaded genomes were simultaneously annotated using Prokka (v1.14.6) [23]. The N5 and N7 strain-specific genes were identified using cd-hit-2d with parameters -s 0.9 and -c 0.9 [24]. The Clusters of Orthologous Groups of proteins (COG) annotations were obtained using emapper (v2.1.7) with default parameters, based on the eggNOG Roary orthology data (v5.0.2) [25], and the functional annotation was also completed by blasting genes against Kyoto Encyclopedia of Genes and Genomes (KEGG) databases [26]. The carbohydrate-active enzyme (CAZyme)-encoding genes in the L. johnsonii genomes were annotated using dbCAN (v2) with default parameters [27]. The average nucleotide identity (ANI) values between the L. johnsonii strains were determined using the recommended methods outlined in FastANI (v1.32) [28]. Roary (v3.13.0) was employed to analyze the pan- and core genome of the 19 L. johnsonii genomes [29].

3. Results and Discussion

3.1. Probiotic Properties of L. johnsonii Strains Isolated from Heat-Resistant Weaned Piglets

In our previous study, we found that L. johnsonii exhibited a relatively high abundance in the microbiota (>15%) of weaned piglets that were resistant to heat stress-induced diarrhea, while it was almost undetectable in the microbiota of those who had diarrhea [19]. The result suggested L. johnsonii probably exhibited probiotic effects to maintain the health of the host. A total of 49 L. johnsonii strains were subsequently isolated from the fecal samples of these heat-resistant piglets, and the beneficial properties of these strains were assessed.
The possession of the capability to inhibit the growth of pathogens is crucial for evaluating the potential probiotic candidacy [30]. We assessed the antagonistic activity of these L. johnsonii strains using the strain S. Typhimurium SL1344, which represents one of the most prominent pathogens in animals, as an indicator. The results demonstrated that the antagonistic ability varied among these L. johnsonii strains (Table S1), and two representative strains, N5 and N7, exhibited superior antagonistic effects compared with others. The tolerance to acidic and bile stresses of strains N5 and N7 were then assessed. This evaluation is pertinent because potential probiotic strains must demonstrate resistance against acidic and bile stresses in the gastrointestinal tract of animals to facilitate colonization. In an acidic environment with a pH of 2.5, the survival rate of L. johnsonii N5 and N7 were 110.34% ± 5.97% (mean ± SD) and 73.33% ± 1.53% after 2 h, and 65.52% ± 6.65% and 33.33% ± 2.25% after 4 h, respectively (Figure 1A). In a 0.3% ox gall solution, the survival rate of L. johnsonii N5 and N7 were 95.74% ± 6.38% (mean ± SD) and 92.31% ± 5.77% after 2 h, and 76.60% ± 6.38% and 113.46% ± 8.81% after 4 h, respectively (Figure 1B). The results demonstrated that the growth activity of N5 and N7 in the acidic and bile environments gradually declined over time, while still maintaining a relatively high survival rate. Then, strain N5 was selected for in vivo probiotic capacity assessment using a mouse model, and the results demonstrated that oral administration of strain N5 alleviated the clinical and histological signs of colitis of the host through several means, including suppressing the expression of pro-inflammatory cytokines TNF-α and IL-6 within the intestinal tract [19].

3.2. General Genomic Features of the L. johnsonii N5 and N7 Strains

To identify potential genomic contents that were associated with the observed probiotic properties as mentioned above, the complete genome sequences of strains L. johnsonii N5 and N7 were determined by a combination of Illumina and nanopore sequencing technologies. A single circular chromosome 1,966,342 bp (Figure 2A) and 1,966,349 bp (Figure 2B) was generated for strains N5 and N7, respectively, with no plasmid identified in either strain. The whole genome ANI value between the two strains was 99.998%, as revealed by FastANI analysis. Then, genes were predicted from the two genomes using PROKKA, and 1863 and 1870 CDSs were identified in N5 and N7. With a threshold of similarity of 90% and coverage of 90% using cd-hit-2d, only two N5-specific and four N7-specific genes were identified, with the remaining genes conserved between the two strains. These results demonstrated that the two strains shared very high genomic contents. The two N5-specific genes included one gene annotated as typA and another annotated as a gene encoding a hypothetical protein, and the four N7-specific genes included two IS30 family transposase ISLga1-encoding genes and two hypothetical protein-encoding genes. TypA encodes a ribosome-binding GTPase that is necessary for survival under some stress conditions and contributes to rapid attachment and biofilm formation [31].
COG and KEGG functional annotations were assigned to the commonly shared genes of the two strains. A total of 1737 CDSs obtained functional annotations that were specifically assigned to 20 COG categories (Figure 3A). The most prevalent COG term was “function unknown” (351 genes, 20.21%), and the remaining genes were primarily classified into functional categories associated with replication, recombination, and repair (190 genes, 10.94%), translation, ribosomal structure and biogenesis (160 genes, 9.21%), carbohydrate transport and metabolism (145 genes, 8.35%), transcription (143 genes, 8.23%), amino acid transport and metabolism (114 genes, 6.56%), cell wall/membrane/envelope biogenesis (99 genes, 5.70%), and nucleotide transport and metabolism (89 genes, 5.12%). Furthermore, a total of 1087 CDSs were assigned to 40 KEGG categories (Figure 3B), mainly functioning in the metabolic pathways (231 genes, 21.25%), biosynthesis of secondary metabolites (86 genes, 7.91%), ABC transporters (53 genes, 4.88%), ribosome (53 genes, 4.88%), microbial metabolism in diverse environments (47 genes, 4.32%), and biosynthesis of cofactors (32 genes, 2.94%).

3.3. Stress Resistance-Associated Genes

Several genes associated with resistance to the acidic and bile stresses were identified in the L. johnsonii N5 and N7 genomes. For instance, cbh, which encodes a chenodeoxycholylglycine hydrolase, and nhaC, which encodes a Na+:H+ antiporter, were identified in both genomes. Cbh and nhaC are found to be essential for the survival of L. johnsonii in acidic and bile salt environments [32]. Two choloylglycine hydrolase-encoding genes were identified in the N5 and N7 genomes. Choloylglycine hydrolase is involved in adaptation to the bile stress and recognized as one of the gut-specific genes of LAB that colonize the intestines [33]. Exopolysaccharides (EPS) produced by LAB play a role in conferring resistance to environmental stresses, including bile salts and acidic pH within the gastrointestinal tract. Additionally, these EPS facilitate the attachment of LAB to intestinal cells in the host, preventing competing pathogenic bacteria from attaching to the host cells [34,35,36]. A generic EPS cluster was identified in strains N5 and N7. Interestingly, in the cluster, epsABCD exhibited very high similarity with the well-studied strain L. johnsonii FI9785 [37] (>90% protein sequence identity for all the four genes); however, epsE exhibited relatively low similarity (protein sequence identity 44.23%) with that of FI9785 (FI9785_1179) and showed high similarity (94.67%) with that of strain MR1, indicative of occurrence of recombination events in the EPS cluster of strains N5 and N7.
Furthermore, several genes associated with resistance to heat stress, including a master transcriptional regulator encoding gene cspA and several genes encoding chaperone-activated proteins, such as HtpX, Hsp20, Hsp33, GroEL, GroES, HslO, and DnaK [38,39], were identified in the L. johnsonii N5 and N7 genomes. Several genes involved in resistance against oxidative stress, such as trxB (encoding a thioredoxin reductase) and msrAB (encoding a peptide–methionine (S)-S-oxide reductase) [38,40], were also identified in the L. johnsonii N5 and N7 genomes. The expression of trxB in L. plantarum WCFS1 significantly enhanced the strain’s resistance to oxidative stress induced by hydrogen peroxide or diamide [41]. HslV, whose product is implicated in stress response associated with the maintenance of protein homeostasis and prevention of protein damage [5], was identified in the L. johnsonii N5 and N7 genomes. Furthermore, the L. johnsonii N5 and N7 genomes were found to harbor 4 sigma factors and 20 transcriptional regulators involved in transcriptional regulation, thereby adequately fulfilling the transcriptional regulatory requirements of L. johnsonii in a nutrient-rich but stressful gut environment. Overall, the presence of these stress resistance-related genes confers enhanced adaptability to gastrointestinal stress conditions in the L. johnsonii N5 and N7 genomes, thereby assisting the improved survival and proliferation of L. johnsonii in response to environmental stresses in the gastrointestinal tract.

3.4. Anti-Inflammation-Associated Genes

Inflammation is a natural response to pathogen infection, but prolonged or uncontrolled inflammation can lead to tissue damage. Our previous study demonstrated that oral administration of L. johnsonii N5 could reduce the expression level of proinflammatory cytokine TNF-α in the intestines [19]. Several anti-inflammation-associated genes were identified in both the N5 and N7 genomes. For instance, groEL and groES were identified in both genomes. The GroEL/GroES chaperone system is responsible for correctly folding proteins in an ATP-regulated manner [42], and GroEL of Lactobacillus has been demonstrated to exhibit an immunomodulatory effect and can inhibit the production of the proinflammatory cytokine TNF-α induced by LPS [43]. Additionally, the N5 and N7 genomes were found to harbor folC, which is involved in folate metabolism and can regulate histamine production. Previous studies suggested that histamine derived from Lactobacillus reuteri can inhibit intestinal inflammation by regulating proinflammatory cytokines through PKA and ERK signaling [44]. Two key genes involved in the biosynthesis of capsular polysaccharides (CPS), capA and capB, were identified in the N5 and N7 genomes. The CPS has been reported to promote macrophage proliferation and the production of regulatory T (Treg) cells in the intestine, which can affect the balance between Tregs and Th17 cells and ultimately prevent the development of colitis [45]. Furthermore, EPS is also believed to have functions in immunomodulation [46]. The presence of the above-mentioned anti-inflammatory genes probably enables L. johnsonii to cope well with intestinal inflammation and protect host health.

3.5. CAZymes

CAZymes are ubiquitously present in various organisms, with a particularly high abundance observed in microorganisms [40,47]. In this study, CAZymes were annotated from L. johnsonii N5 and N7 genomes using dbCAN2, and an identical number of CAZymes were identified in N5 and N7 genomes, including three carbohydrate esterase (CE) families, 13 carbohydrate-binding modules (CBM) families, 57 glycosyl transferases (GT) families, and 62 glycoside hydrolases (GH) families (Figure 4). Genes belonging to GHs and GTs exhibited the highest number in L. johnsonii N5 and N7, which is consistent with that identified in other Lactobacillus species including L. salivarius [17]. The GHs play a crucial role in the degradation of glycosidic bonds between the chain residues of the carbohydrate compounds, with GH13 being the predominant subgroup in L. johnsonii N5 and N7. GH13 is essential for the degradation of various carbohydrates including amylase, β-glucosidase, β-xylosidase, and β-galactosidase [48]. GTs facilitate the formation of glycosidic bonds by transferring sugar residues from donors to acceptors, encompassing carbohydrates, proteins, lipids, DNA, and other biomolecules [49,50]. GT2 and GT4 were the most predominant GT subgroups in L. johnsonii N5 and N7. Generally, GT2 is responsible for the biosynthesis of cellulose, undecaprenyl-diphospho-muramoyl pentapeptide, beta-N-acetylglucosaminyltransferase, and putative glycosyltransferase, while GT4 is responsible for facilitating the activity of the Sec-dependent glycosyltransferase (gtfA), which plays a crucial role in sucrose biosynthesis [51,52]. Overall, these genes actively participate in the metabolism and transportation of functional and bioactive substances, thereby facilitating efficient nutrient transport and synthesis by L. johnsonii N5 and N7 within the intestinal tract.

3.6. Biosynthesis and Transport System

Both N5 and N7 harbored asnA and asnB, encoding asparagine synthetases. Asparagine synthetases are involved in the synthesis of L-asparagine from L-aspartic acid, contributing to efficient ammonia assimilation [53]. Additionally, both strains also carried glnA, whose product is a glutamine synthetase (GLUL) that catalyzes the formation of glutamine by combining glutamic acid and ammonia. The presence of glycine hydroxymethyltransferase (SHMT) encoded by glyA suggested that L. johnsonii N5 and N7 can facilitate the interconversion between serine and glycine. Moreover, L. johnsonii N5 and N7 could synthesize lysine from aspartic acid through a series of enzymes encoded by lysC, asd, dapA, dapB, dapH, patA, dapL, dapF, and lysA. Furthermore, L. johnsonii N5 and N7 exhibited robust transport systems, as evidenced by the presence of a total of 70 genes associated with transport mechanisms in their genomes. Notably, these included 53 ABC transport system genes that are responsible for the efficient transportation of amino acids and peptides, thereby facilitating protein synthesis (Table S2). The genomes of L. johnsonii N5 and N7 also contained 17 genes associated with the phosphoenolpyruvate-dependent sugar phosphotransferase system (PTS) including PTS enzyme I (PTS-EI) and PTS-EII complexes (Table S2), which catalyze the phosphorylation of carbohydrate substrates across the microbial cell membrane and serve as a prominent active carbohydrate transport mechanism in bacteria [54]. Specifically, the L. johnsonii N5 and N7 genomes carried ptsI and ptsH, which encoded PTS-EI and the phosphor carrier protein HPR, respectively. The two genes collaborate to facilitate the transfer and transport of phosphorylated compounds by transferring phosphoenolpyruvate to PTS-EII [5,38]. The genomes of N5 and N7 encoded 14 PTS-EII complexes, surpassing the count observed in other microorganisms with comparable genome sizes [53]. The information of these complexes is closely linked to the transport of diverse carbon sources, including glucose, β-glucosides, disaccharides, mannitol, sorbitol, mannose, and sucrose [55]. To validate the predictions on carbohydrate metabolism, we conducted Analytical Profile Index (API) identification using a rapid API-50 CH biochemical test kit. The biochemical activities of D-galactose, D-glucose, salicin, D-cellobiose, D-maltose, D-melibiose, D-sucrose, D-trehalose, mannitol, methyl-β-D-fucopyranoside, and methyl-α-D-glucopyranoside are positive (Table 2), further suggesting that these PTSs enable the introduction of multiple substrates and facilitate the transport of diverse sugars, thus augmenting the carbon transport capacity of L. johnsonii N5 and N7 genomes.

3.7. Genome Comparison of Strains N5 and N7 and Related Complete L. johnsonii Genomes

Comparative genomic analyses were performed using our newly sequenced N5 and N7 genomes and 17 complete L. johnsonii genomes available in the NCBI Refseq database. ANI was used to assess genomic relatedness between these L. johnsonii strains [56,57]. The commonly employed threshold of ANI values for demarcating bacterial species based on their genomic sequence is 95% [28]. In this study, the ANI values of the 19 L. johnsonii strains were all higher than 95% (Figure S1), indicating that these strains were affiliated with the same species. A phylogenetic tree based on the core genome sequences was constructed to enhance our understanding of the phylogenetic relationships among strains of L. johnsonii. The phylogenetic tree was classified into three distinct clusters (Figure 5), and L. johnsonii N5 and N7 were found to be closely associated with GHZ10a and ZLJ010, which were isolated from swine feces [5].
The core genome and pan-genome are commonly employed for evaluating the genomic diversity of a species or closely related bacteria [58]. The core genome of a species encompasses the complete set of genes shared by all strains, including the genetic determinants that maintain species-specific characteristics; on the other hand, the pan-genome represents the collective gene pool that can be accommodated by the genetic determinants, including core genes and accessory genes (i.e., genes not harbored by all strains) [59,60]. The pan-genome of L. johnsonii comprised 5977 genes across 19 strains, and its expansion with the addition of newly sequenced genomes signified an open nature of the L. johnsonii pan-genome (Figure 6A). The estimated number of newly added genes per genome sequence revealed a rapid decline in the average count from the second genome sequence onwards, interspersed with occasional fluctuations indicating sporadic increases in subsequent genomes (Figure 6B). Even in the 19th genome, novel genes continued to be incorporated, suggesting ongoing gene exchange within L. johnsonii species and between other bacterial species, with a predominant emergence of new functionalities during evolutionary processes. This observation may be attributed to the extensive genetic variability among strains, enabling them to effectively adapt to the host intestinal environment, and it is plausible that these phenomena could also be associated with the insertion and translocation of insertion sequences (IS), thereby facilitating gene recombination [61]. These accessory genes could contribute to bacterial evolution and adaptation to diverse environments [62]. In contrast to the pan-genome, the core genome of L. johnsonii contained 1044 genes and exhibited gradual stabilization as the number of genomes increased to 15. Consequently, an increase in the number of genomes exerted a relatively minor influence on the magnitude of the core genome.

3.8. Unique Genomic Characteristics of L. johnsonii N5 and N7

The L. johnsonii N5 and N7 genomes collectively harbor a total of 115 unique genes, compared with the other 17 complete L. johnsonii genomes (Table S3), and 99 out of these 115 genes showed homology with genes deposited in the NCBI nr database; 16 genes were identified as novel genes based on the BLAST analysis (identity threshold 80% and coverage threshold 80%). The potential function of these N5- and N7-specifically harbored genes was then predicted, 86 genes obtained COG annotation and 29 genes were identified as hypothetical proteins with undetermined functions. The 86 COG annotated genes were assigned to 17 COG categories, with “replication, recombination, and repair” (L) being the most prevalent category (including 45 genes, accounting for 52.3% of the 86 annotated genes). Among the 45 genes assigned to COG category L, 37 genes encode transposases, and 20 of these encode transposases belonging to the IS30 family, which can enable the insertion or removal of DNA fragments at new positions in the genomes [40,63]. The process of gene recombination can enhance the diversity and adaptability of L. johnsonii in diverse environments, thereby conferring advantageous traits for survival and reproduction within the host’s surrounding milieu [63]. Furthermore, nine genes belonged to the “transcription” category (K), including several transcriptional regulators. Among them, one gene was annotated to encode a transcriptional factor AsnR. AsnR plays an important role in regulating the utilization of asparagine by the bacterial host [64]. A RimI encoding gene was also identified in the genes belonging to COG category K. RimI is demonstrated to be important in regulating the growth rate of bacterial cells by modification of bacterial translation apparatus [65]. These regulators possibly exert their influence on specific genes, thereby governing gene expression and fulfilling crucial regulatory functions in bacterial metabolism, stress response, and other physiological processes, enabling the strain to acquire advantageous traits within the intestinal tract [66]. The “cell wall/membrane/envelope biogenesis” category (M) encompassed nine genes, primarily encoding GTs, and these GTs are closely associated with EPS synthesis [67]. Previous research has demonstrated that GTs, particularly glucosyltransferase (GTF), are essential for EPS biosynthesis [68,69]. Therefore, it is postulated that the presence of these specific GTs confers a competitive advantage to L. johnsonii N5 and N7 in the host’s intestinal environment. Of note, a large fraction of the N5- and N7-specifically harbored genes lacked definite functional annotations (Table S3). The functions of these genes need to be further determined by future experiments, such as gene deletion-, transcriptomics-, and proteomics-based approaches.

4. Conclusions

Considering the plethora of probiotic properties of L. johnsonii N5 and N7, these strains could be considered as potential probiotic candidates. The complete genome sequences of the two strains were obtained by the integrative combination of short-read Illumina and long-read nanopore sequencing. The two strains harbored highly conserved genomic backgrounds, with two N5-specific and four N7-specific genes identified. Several genes that were involved in adaptation in the gastrointestinal environments and probiotic properties including anti-inflammation were identified in both genomes. Through comparative genomics analysis with 17 available complete L. johnsonii genomes, 115 genes that were specifically harbored by N5 and N7 were identified, which mainly encoded proteins involved in transposase, replication, recombination and repair, and transcription. The presence of genes with probiotic properties such as those involved in resistance against acidic and bile stresses, anti-inflammation, and CAZymes, probably enables N5 and N7 to efficiently adapt to the host environment and exert beneficial effects. In summary, the genome-wide analysis of the two L. johnsonii strains provides further insight into the niche adaptation and probiotic properties of L. johnsonii and provides clues for targeted experimental studies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/microorganisms12040672/s1, Figure S1: Heatmap of ANI based on the sequences of 19 Lactobacilli johnsonii; Table S1: Antagonistic activity of Lactobacilli johnsonii strains against Salmonella Typhimurium SL1344; Table S2: Summary of genes encoding ABC transporters and PTS systems in strains N5 and N7; Table S3: COG annotations of genes specifically harbored by strains N5 and N7.

Author Contributions

Conceptualization, Y.W. and Y.Z.; Formal analysis, K.W., Y.W. and L.G.; Funding acquisition, X.J.; Investigation, F.G. and H.L.; Methodology, K.W., Y.W., L.G., J.Y. and Q.L.; Project administration, Y.Z.; Resources, J.Y., Q.L., R.Z., G.L., H.C., F.G. and H.L.; Supervision, X.J.; Visualization, G.L.; Writing—original draft, K.W.; Writing—review and editing, Y.W. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by funded by the 111 Project (grant number D18007), the Priority Academic Program Development of Jiangsu Higher Education Institutions (grant number PAPD), and the Postgraduate Research and Practice Innovation Program of Jiangsu Province (grant number SJCX22_1769).

Data Availability Statement

The genomic sequences of L. johnsonii N5 and N7 have been deposited at GenBank with the accession numbers CP136013 and CP136012.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gómez, N.C.; Ramiro, J.M.; Quecan, B.X.; de Melo Franco, B.D. Use of Potential Probiotic Lactic Acid Bacteria (LAB) Biofilms for the Control of Listeria monocytogenes, Salmonella Typhimurium, and Escherichia coli O157:H7 Biofilms Formation. Front. Microbiol. 2016, 7, 863. [Google Scholar] [CrossRef]
  2. Camargo, A.C.; Todorov, S.D.; Chihib, N.E.; Drider, D.; Nero, L.A. Lactic Acid Bacteria (LAB) and Their Bacteriocins as Alternative Biotechnological Tools to Control Listeria monocytogenes Biofilms in Food Processing Facilities. Mol. Biotechnol. 2018, 60, 712–726. [Google Scholar] [CrossRef] [PubMed]
  3. Miki, T.; Goto, R.; Fujimoto, M.; Okada, N.; Hardt, W.-D. The Bactericidal Lectin RegIIIβ Prolongs Gut Colonization and Enteropathy in the Streptomycin Mouse Model for Salmonella Diarrhea. Cell Host Microbe 2017, 21, 195–207. [Google Scholar] [CrossRef] [PubMed]
  4. Kant, R.; Rintahaka, J.; Yu, X.; Sigvart-Mattila, P.; Paulin, L.; Mecklin, J.-P.; Saarela, M.; Palva, A.; von Ossowski, I. A Comparative Pan-Genome Perspective of Niche-Adaptable Cell-Surface Protein Phenotypes in Lactobacillus rhamnosus. PLoS ONE 2014, 9, e102762. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, W.; Wang, J.; Zhang, D.; Liu, H.; Wang, S.; Wang, Y.; Ji, H. Complete Genome Sequencing and Comparative Genome Characterization of Lactobacillus johnsonii ZLJ010, a Potential Probiotic with Health-Promoting Properties. Front. Genet. 2019, 10, 812. [Google Scholar] [CrossRef] [PubMed]
  6. Masuoka, H.; Shimada, K.; Kiyosue-Yasuda, T.; Kiyosue, M.; Oishi, Y.; Kimura, S.; Yamada, A.; Hirayama, K. Transition of the intestinal microbiota of dogs with age. Biosci. Microbiota Food Health 2017, 36, 27–31. [Google Scholar] [CrossRef] [PubMed]
  7. Johnson, A.; Miller, E.A.; Weber, B.; Figueroa, C.F.; Aguayo, J.M.; Johny, A.K.; Noll, S.; Brannon, J.; Kozlowicz, B.; Johnson, T.J. Evidence of host specificity in Lactobacillus johnsonii genomes and its influence on probiotic potential in poultry. Poult. Sci. 2023, 102, 102858. [Google Scholar] [CrossRef]
  8. Li, Y.; Li, X.; Nie, C.; Wu, Y.; Luo, R.; Chen, C.; Niu, J.; Zhang, W. Effects of two strains of Lactobacillus isolated from the feces of calves after fecal microbiota transplantation on growth performance, immune capacity, and intestinal barrier function of weaned calves. Front. Microbiol. 2023, 14, 1249628. [Google Scholar] [CrossRef] [PubMed]
  9. Kim, D.; Cho, M.-J.; Cho, S.; Lee, Y.; Byun, S.J.; Lee, S. Complete Genome Sequence of Lactobacillus johnsonii Strain Byun-jo-01, Isolated from the Murine Gastrointestinal Tract. Microbiol. Resour. Announc. 2018, 7, e00985-18. [Google Scholar] [CrossRef]
  10. Audisio, M.C.; Benítez-Ahrendts, M.R. Lactobacillus johnsonii CRL1647, isolated from Apis mellifera L. bee-gut, exhibited a beneficial effect on honeybee colonies. Benefic. Microbes 2011, 2, 29–34. [Google Scholar] [CrossRef]
  11. Zhang, X.; Mushajiang, S.; Luo, B.; Tian, F.; Ni, Y.; Yan, W. The Composition and Concordance of Lactobacillus Populations of Infant Gut and the Corresponding Breast-Milk and Maternal Gut. Front. Microbiol. 2020, 11, 597911. [Google Scholar] [CrossRef] [PubMed]
  12. Salvetti, E.; O’toole, P.W. The Genomic Basis of Lactobacilli as Health-Promoting Organisms. Microbiol. Spectr. 2017, 5, 3. [Google Scholar] [CrossRef] [PubMed]
  13. Koutsoumanis, K.; Allende, A.; Alvarez-Ordóñez, A.; Bolton, D.; Bover-Cid, S.; Chemaly, M.; Davies, R.; De Cesare, A.; Hilbert, F.; Lindqvist, R.; et al. Update of the list of QPS-recommended biological agents intentionally added to food or feed as notified to EFSA 15: Suitability of taxonomic units notified to EFSA until September 2021. EFSA J. 2022, 20, e07045. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, H.; Ni, X.; Qing, X.; Zeng, D.; Luo, M.; Liu, L.; Li, G.; Pan, K.; Jing, B. Live Probiotic Lactobacillus johnsonii BS15 Promotes Growth Performance and Lowers Fat Deposition by Improving Lipid Metabolism, Intestinal Development, and Gut Microflora in Broilers. Front. Microbiol. 2017, 8, 1073. [Google Scholar] [CrossRef] [PubMed]
  15. Yin, Z.; Wang, K.; Liu, Y.; Li, Y.; He, F.; Yin, J.; Tang, W. Lactobacillus johnsonii Improves Intestinal Barrier Function and Reduces Post-Weaning Diarrhea in Piglets: Involvement of the Endocannabinoid System. Animals 2024, 14, 493. [Google Scholar] [CrossRef] [PubMed]
  16. Tedersoo, L.; Albertsen, M.; Anslan, S.; Callahan, B. Perspectives and Benefits of High-Throughput Long-Read Sequencing in Microbial Ecology. Appl. Environ. Microbiol. 2021, 87, e00626-21. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, Y.; Xu, X.; Chen, H.; Yang, F.; Xu, B.; Wang, K.; Liu, Q.; Liang, G.; Zhang, R.; Jiao, X.; et al. Assessment of beneficial effects and identification of host adaptation-associated genes of Ligilactobacillus salivarius isolated from badgers. BMC Genom. 2023, 24, 530. [Google Scholar] [CrossRef] [PubMed]
  18. Mendoza, R.M.; Kim, S.H.; Vasquez, R.; Hwang, I.-C.; Park, Y.-S.; Paik, H.-D.; Moon, G.-S.; Kang, D.-K. Bioinformatics and its role in the study of the evolution and probiotic potential of lactic acid bacteria. Food Sci. Biotechnol. 2023, 32, 389–412. [Google Scholar] [CrossRef]
  19. Yuan, L.; Zhu, C.; Gu, F.; Zhu, M.; Yao, J.; Zhu, C.; Li, S.; Wang, K.; Hu, P.; Zhang, Y.; et al. Lactobacillus johnsonii N5 from heat stress-resistant pigs improves gut mucosal immunity and barrier in dextran sodium sulfate-induced colitis. Anim. Nutr. 2023, 15, 210–224. [Google Scholar] [CrossRef]
  20. Leite, A.M.; Miguel, M.A.; Peixoto, R.S.; Ruas-Madiedo, P.; Paschoalin, V.M.; Mayo, B.; Delgado, S. Probiotic potential of selected lactic acid bacteria strains isolated from Brazilian kefir grains. J. Dairy Sci. 2015, 98, 3622–3632. [Google Scholar] [CrossRef]
  21. Kolmogorov, M.; Yuan, J.; Lin, Y.; Pevzner, P.A. Assembly of long, error-prone reads using repeat graphs. Nat. Biotechnol. 2019, 37, 540–546. [Google Scholar] [CrossRef]
  22. Hu, J.; Fan, J.; Sun, Z.; Liu, S. NextPolish: A fast and efficient genome polishing tool for long-read assembly. Bioinformatics 2020, 36, 2253–2255. [Google Scholar] [CrossRef] [PubMed]
  23. Seemann, T. Prokka: Rapid Prokaryotic Genome Annotation. Bioinformatics 2014, 30, 2068–2069. [Google Scholar] [CrossRef]
  24. Huang, Y.; Niu, B.; Gao, Y.; Fu, L.; Li, W. CD-HIT Suite: A web server for clustering and comparing biological sequences. Bioinformatics 2010, 26, 680–682. [Google Scholar] [CrossRef] [PubMed]
  25. Cantalapiedra, C.P.; Hernández-Plaza, A.; Letunic, I.; Bork, P.; Huerta-Cepas, J. eggNOG-mapper v2: Functional Annotation, Orthology Assignments, and Domain Prediction at the Metagenomic Scale. Mol. Biol. Evol. 2021, 38, 5825–5829. [Google Scholar] [CrossRef] [PubMed]
  26. Kanehisa, M.; Goto, S. KEGG: Kyoto Encyclopedia of Genes and Genomes. Nucleic Acids Res. 2000, 28, 27–30. [Google Scholar] [CrossRef] [PubMed]
  27. Yin, Y.; Mao, X.; Yang, J.; Chen, X.; Mao, F.; Xu, Y. dbCAN: A web resource for automated carbohydrate-active enzyme an-notation. Nucleic Acids Res. 2012, 40, W445–W451. [Google Scholar] [CrossRef] [PubMed]
  28. Jain, C.; Rodriguez-R, L.M.; Phillippy, A.M.; Konstantinidis, K.T.; Aluru, S. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat. Commun. 2018, 9, 5114. [Google Scholar] [CrossRef] [PubMed]
  29. Page, A.J.; Cummins, C.A.; Hunt, M.; Wong, V.K.; Reuter, S.; Holden, M.T.; Fookes, M.; Falush, D.; Keane, J.A.; Parkhill, J. Roary: Rapid large-scale prokaryote pan genome analysis. Bioinformatics 2015, 31, 3691–3693. [Google Scholar] [CrossRef]
  30. Liu, J.; Hu, D.; Chen, Y.; Huang, H.; Zhang, H.; Zhao, J.; Gu, Z.; Chen, W. Strain-specific properties of Lactobacillus plantarum for prevention of Salmonella infection. Food Funct. 2018, 9, 3673–3682. [Google Scholar] [CrossRef]
  31. Neidig, A.; Yeung, A.T.; Rosay, T.; Tettmann, B.; Strempel, N.; Rueger, M.; Lesouhaitier, O.; Overhage, J. TypA is involved in virulence, antimicrobial resistance and biofilm formation in Pseudomonas aeruginosa. BMC Microbiol. 2013, 13, 77. [Google Scholar] [CrossRef] [PubMed]
  32. Xu, S.; Cheng, J.; Meng, X.; Xu, Y.; Mu, Y. Complete Genome and Comparative Genome Analysis of Lactobacillus reuteri YSJL-12, a Potential Probiotics Strain Isolated from Healthy Sow Fresh Feces. Evol. Bioinform. 2020, 16, 1176934320942192. [Google Scholar] [CrossRef]
  33. O’Sullivan, O.; O’Callaghan, J.; Sangrador-Vegas, A.; McAuliffe, O.; Slattery, L.; Kaleta, P.; Callanan, M.; Fitzgerald, G.F.; Ross, R.P.; Beresford, T. Comparative genomics of lactic acid bacteria reveals a niche-specific gene set. BMC Microbiol. 2009, 9, 50. [Google Scholar] [CrossRef]
  34. Dertli, E.; Mayer, M.J.; Narbad, A. Impact of the exopolysaccharide layer on biofilms, adhesion and resistance to stress in Lactobacillus johnsonii FI9785. BMC Microbiol. 2015, 15, 8. [Google Scholar] [CrossRef] [PubMed]
  35. Werning, M.L.; Hernández-Alcántara, A.M.; Ruiz, M.J.; Soto, L.P.; Dueñas, M.T.; López, P.; Frizzo, L.S. Biological Functions of Exopolysaccharides from Lactic Acid Bacteria and Their Potential Benefits for Humans and Farmed Animals. Foods 2022, 11, 1284. [Google Scholar] [CrossRef]
  36. Ruas-Madiedo, P.; Gueimonde, M.; Margolles, A.; de los Reyes-Gavilán, C.G.; Salminen, S. Exopolysaccharides produced by probiotic strains modify the adhesion of probiotics and enteropathogens to human intestinal mucus. J. Food Prot. 2006, 69, 2011–2015. [Google Scholar] [CrossRef]
  37. Dertli, E.; Colquhoun, I.J.; Gunning, A.P.; Bongaerts, R.J.; Le Gall, G.; Bonev, B.B.; Mayer, M.J.; Narbad, A. Structure and Biosynthesis of Two Exopolysaccharides Produced by Lactobacillus johnsonii FI9785. J. Biol. Chem. 2013, 288, 31938–31951. [Google Scholar] [CrossRef] [PubMed]
  38. Boucard, A.-S.; Florent, I.; Polack, B.; Langella, P.; Bermúdez-Humarán, L.G. Genome Sequence and Assessment of Safety and Potential Probiotic Traits of Lactobacillus johnsonii CNCM I-4884. Microorganisms 2022, 10, 273. [Google Scholar] [CrossRef]
  39. Papadimitriou, K.; Alegría, Á.; Bron, P.A.; de Angelis, M.; Gobbetti, M.; Kleerebezem, M.; Lemos, J.A.; Linares, D.M.; Ross, P.; Stanton, C.; et al. Stress Physiology of Lactic Acid Bacteria. Microbiol. Mol. Biol. Rev. 2016, 80, 837–890. [Google Scholar] [CrossRef]
  40. Chen, L.; Gu, Q.; Li, P.; Chen, S.; Li, Y. Genomic analysis of Lactobacillus reuteri WHH1689 reveals its probiotic properties and stress resistance. Food Sci. Nutr. 2019, 7, 844–857. [Google Scholar] [CrossRef]
  41. Serrano, L.M.; Molenaar, D.; Wels, M.; Teusink, B.; A Bron, P.; de Vos, W.M.; Smid, E.J. Thioredoxin reductase is a key factor in the oxidative stress response of Lactobacillus plantarum WCFS1. Microb. Cell Factories 2007, 6, 29. [Google Scholar] [CrossRef] [PubMed]
  42. Aoudia, N.; Rieu, A.; Briandet, R.; Deschamps, J.; Chluba, J.; Jego, G.; Garrido, C.; Guzzo, J. Biofilms of Lactobacillus plantarum and Lactobacillus fermentum: Effect on stress responses, antagonistic effects on pathogen growth and immunomodulatory properties. Food Microbiol. 2016, 53, 51–59. [Google Scholar] [CrossRef] [PubMed]
  43. Rieu, A.; Aoudia, N.; Jego, G.; Chluba, J.; Yousfi, N.; Briandet, R.; Deschamps, J.; Gasquet, B.; Monedero, V.; Garrido, C.; et al. The biofilm mode of life boosts the anti-inflammatory properties of Lactobacillus. Cell. Microbiol. 2014, 16, 1836–1853. [Google Scholar] [CrossRef] [PubMed]
  44. Thomas, C.M.; Hong, T.; Van Pijkeren, J.P.; Hemarajata, P.; Trinh, D.V.; Hu, W.; Britton, R.A.; Kalkum, M.; Versalovic, J. Histamine Derived from Probiotic Lactobacillus reuteri Suppresses TNF via Modulation of PKA and ERK Signaling. PLoS ONE 2012, 7, e31951. [Google Scholar] [CrossRef] [PubMed]
  45. Li, J.; Feng, S.; Yu, L.; Zhao, J.; Tian, F.; Chen, W.; Zhai, Q. Capsular polysaccarides of probiotics and their immunomodulatory roles. Food Sci. Hum. Wellness 2022, 11, 1111–1120. [Google Scholar] [CrossRef]
  46. Yasuda, E.; Serata, M.; Sako, T. Suppressive Effect on Activation of Macrophages by Lactobacillus casei Strain Shirota Genes Determining the Synthesis of Cell Wall-Associated Polysaccharides. Appl. Environ. Microbiol. 2008, 74, 4746–4755. [Google Scholar] [CrossRef] [PubMed]
  47. El Kaoutari, A.; Armougom, F.; Gordon, J.I.; Raoult, D.; Henrissat, B. The abundance and variety of carbohydrate-active enzymes in the human gut microbiota. Nat. Rev. Microbiol. 2013, 11, 497–504. [Google Scholar] [CrossRef]
  48. Janeček, Š.; Svensson, B.; MacGregor, E.A. α-Amylase: An enzyme specificity found in various families of glycoside hydrolases. Cell. Mol. Life Sci. 2014, 71, 1149–1170. [Google Scholar] [CrossRef] [PubMed]
  49. Costa, O.Y.A.; de Hollander, M.; Pijl, A.; Liu, B.; Kuramae, E.E. Cultivation-independent and cultivation-dependent metagenomes reveal genetic and enzymatic potential of microbial community involved in the degradation of a complex microbial polymer. Microbiome 2020, 8, 76. [Google Scholar] [CrossRef]
  50. Ardèvol, A.; Rovira, C. Reaction Mechanisms in Carbohydrate-Active Enzymes: Glycoside Hydrolases and Glycosyltransferases. Insights from ab Initio Quantum Mechanics/Molecular Mechanics Dynamic Simulations. J. Am. Chem. Soc. 2015, 137, 7528–7547. [Google Scholar] [CrossRef]
  51. Kumar, S.; Bansal, K.; Sethi, S.K.; Bindhani, B.K. Phylo-taxonogenomics of 182 strains of genus Leuconostoc elucidates its robust taxonomy and biotechnological importance. BioRxiv 2021. [Google Scholar] [CrossRef]
  52. Mao, B.; Yin, R.; Li, X.; Cui, S.; Zhang, H.; Zhao, J.; Chen, W. Comparative Genomic Analysis of Lactiplantibacillus plantarum Isolated from Different Niches. Genes 2021, 12, 241. [Google Scholar] [CrossRef] [PubMed]
  53. Pridmore, R.D.; Berger, B.; Desiere, F.; Vilanova, D.; Barretto, C.; Pittet, A.-C.; Zwahlen, M.-C.; Rouvet, M.; Altermann, E.; Barrangou, R.; et al. The genome sequence of the probiotic intestinal bacterium Lactobacillus johnsonii NCC 533. Proc. Natl. Acad. Sci. USA 2004, 101, 2512–2517. [Google Scholar] [CrossRef] [PubMed]
  54. Deutscher, J.; Francke, C.; Postma, P.W. How Phosphotransferase System-Related Protein Phosphorylation Regulates Carbohydrate Metabolism in Bacteria. Microbiol. Mol. Biol. Rev. 2006, 70, 939–1031. [Google Scholar] [CrossRef]
  55. Liu, C.-J.; Wang, R.; Gong, F.-M.; Liu, X.-F.; Zheng, H.-J.; Luo, Y.-Y.; Li, X.-R. Complete genome sequences and comparative genome analysis of Lactobacillus plantarum strain 5-2 isolated from fermented soybean. Genomics 2015, 106, 404–411. [Google Scholar] [CrossRef] [PubMed]
  56. Richter, M.; Rosselló-Móra, R. Shifting the genomic gold standard for the prokaryotic species definition. Proc. Natl. Acad. Sci. USA 2009, 106, 19126–19131. [Google Scholar] [CrossRef]
  57. Kim, M.; Oh, H.-S.; Park, S.-C.; Chun, J. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 2014, 64, 346–351. [Google Scholar] [CrossRef]
  58. Medini, D.; Donati, C.; Tettelin, H.; Masignani, V.; Rappuoli, R. The microbial pan-genome. Curr. Opin. Genet. Dev. 2005, 15, 589–594. [Google Scholar] [CrossRef] [PubMed]
  59. Tettelin, H.; Riley, D.; Cattuto, C.; Medini, D. Comparative genomics: The bacterial pan-genome. Curr. Opin. Microbiol. 2008, 11, 472–477. [Google Scholar] [CrossRef]
  60. Huang, Z.; Zhou, X.; Stanton, C.; Ross, R.P.; Zhao, J.; Zhang, H.; Yang, B.; Chen, W. Comparative Genomics and Specific Functional Characteristics Analysis of Lactobacillus acidophilus. Microorganisms 2021, 9, 1992. [Google Scholar] [CrossRef]
  61. Guinane, C.M.; Kent, R.M.; Norberg, S.; Hill, C.; Fitzgerald, G.F.; Stanton, C.; Ross, R.P. Host Specific Diversity in Lactobacillus johnsonii as Evidenced by a Major Chromosomal Inversion and Phage Resistance Mechanisms. PLoS ONE 2011, 6, e18740. [Google Scholar] [CrossRef] [PubMed]
  62. Ventura, M.; Canchaya, C.; Pridmore, R.D.; Brüssow, H. The prophages of Lactobacillus johnsonii NCC 533: Comparative genomics and transcription analysis. Virology 2004, 320, 229–242. [Google Scholar] [CrossRef] [PubMed]
  63. Kumar, R.; Grover, S.; Kaushik, J.K.; Batish, V.K. IS30-related transposon mediated insertional inactivation of bile salt hydrolase (bsh1) gene of Lactobacillus plantarum strain Lp20. Microbiol. Res. 2014, 169, 553–560. [Google Scholar] [CrossRef] [PubMed]
  64. Fisher, S.H.; Wray, L.V., Jr. Bacillus subtilis 168 contains two differentially regulated genes encoding L-asparaginase. J. Bacteriol. 2002, 184, 2148–2154. [Google Scholar] [CrossRef] [PubMed]
  65. Pletnev, P.I.; Shulenina, O.; Evfratov, S.; Treshin, V.; Subach, M.F.; Serebryakova, M.V.; Osterman, I.A.; Paleskava, A.; Bogdanov, A.A.; Dontsova, O.A.; et al. Ribosomal protein S18 acetyltransferase RimI is responsible for the acetylation of elongation factor Tu. J. Biol. Chem. 2022, 298, 101914. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, H.; Zhang, Y.; Li, L.; Li, Y.; Yuan, L.; E, Y.; Qiao, J. Positive regulation of the DLT operon by TCSR7 enhances acid tolerance of Lactococcus lactis F44. J. Dairy Sci. 2022, 105, 7940–7950. [Google Scholar] [CrossRef] [PubMed]
  67. Jia, Y.; Yang, B.; Ross, P.; Stanton, C.; Zhang, H.; Zhao, J.; Chen, W. Comparative Genomics Analysis of Lactobacillus mucosae from Different Niches. Genes 2020, 11, 95. [Google Scholar] [CrossRef] [PubMed]
  68. Mayer, M.J.; D’amato, A.; Colquhoun, I.J.; Le Gall, G.; Narbad, A. Identification of Genes Required for Glucan Exopolysaccharide Production in Lactobacillus johnsonii Suggests a Novel Biosynthesis Mechanism. Appl. Environ. Microbiol. 2020, 86, e02808-19. [Google Scholar] [CrossRef]
  69. Altermann, E.; Klaenhammer, T.R. Group-specific comparison of four lactobacilli isolated from human sources using differential blast analysis. Genes Nutr. 2011, 6, 319–340. [Google Scholar] [CrossRef]
Figure 1. Survival rates of L. johnsonii N5 and N7 under pH 2.5 (A) and bile salt (B) stresses. Data shown are mean ± SD of triplicate values of independent experiments.
Figure 1. Survival rates of L. johnsonii N5 and N7 under pH 2.5 (A) and bile salt (B) stresses. Data shown are mean ± SD of triplicate values of independent experiments.
Microorganisms 12 00672 g001
Figure 2. Genome atlases of L. johnsonii N5 (A) and L. johnsonii N7 (B). Each side of the figure consists of five circles. From the outermost circle to the inside, circles 1 and 2 show the distribution of coding genes in the front chain and the back chain, including tRNA (purple) and rRNA (green), circle 3 shows the GC content, and circles 4 and 5 show GC Skew+ and GC Skew- respectively.
Figure 2. Genome atlases of L. johnsonii N5 (A) and L. johnsonii N7 (B). Each side of the figure consists of five circles. From the outermost circle to the inside, circles 1 and 2 show the distribution of coding genes in the front chain and the back chain, including tRNA (purple) and rRNA (green), circle 3 shows the GC content, and circles 4 and 5 show GC Skew+ and GC Skew- respectively.
Microorganisms 12 00672 g002
Figure 3. The number of genes assigned to COG (A) and KEGG (B) categories. COG, Clusters of Orthologous Group; KEGG, Kyoto Encyclopedia of Genes and Genomes.
Figure 3. The number of genes assigned to COG (A) and KEGG (B) categories. COG, Clusters of Orthologous Group; KEGG, Kyoto Encyclopedia of Genes and Genomes.
Microorganisms 12 00672 g003
Figure 4. Distribution of CAZymes in the L. johnsonii N5 and N7 genomes.
Figure 4. Distribution of CAZymes in the L. johnsonii N5 and N7 genomes.
Microorganisms 12 00672 g004
Figure 5. Phylogenetic relationship between representative L. johnsonii genomes and strains N5 and N7 as revealed by Roary analysis, and displayed using iTOL. The genome size, origin, and number of CDSs are listed alongside the strain name.
Figure 5. Phylogenetic relationship between representative L. johnsonii genomes and strains N5 and N7 as revealed by Roary analysis, and displayed using iTOL. The genome size, origin, and number of CDSs are listed alongside the strain name.
Microorganisms 12 00672 g005
Figure 6. L. johnsonii N5 and N7 pan-genome (A). Number of new genes identified in L. johnsonii with sequential addition of new genomes (B).
Figure 6. L. johnsonii N5 and N7 pan-genome (A). Number of new genes identified in L. johnsonii with sequential addition of new genomes (B).
Microorganisms 12 00672 g006
Table 1. Summary of Lactobacillus johnsonii genomes used in this study.
Table 1. Summary of Lactobacillus johnsonii genomes used in this study.
StrainRefSeq Assembly AccessionGC%Size (Mb)CDSPlasmidIsolation SourceHostCountry
N5 (this study)GCF_032463685.1351.8818630FecesSwineChina
N7 (this study)GCF_032463545.1351.8818700FecesSwineChina
GHZ10aGCF_014841035.134.92.0119212FecesSwineChina
7409N31GCF_022810665.1352.1921460FecesBovineNA
BS15GCF_001714745.134.92.1620761Yoghurt-China
ZLJ010GCF_004011315.134.92.0019590IntestineSwineChina
UMNLJ22GCF_002176835.134.61.9919222IleumPoultryUSA
UMNLJ21GCF_002176855.135.122.0118882IleumPoultryUSA
Byun-jo-01GCF_003316915.134.71.9617810JejunumMouseSouth Korea
MR1GCF_022213385.134.71.9518050Mice cecumMouseUSA
NCK2677GCF_014058685.134.71.9517920MouseMouseUSA
G2AGCF_010586925.134.62.1720452MouseMouseUSA
DC22.2GCF_009769185.134.51.9418584BirdsJunglefowlUK
IDCC9203GCF_003428395.134.71.9018820Infant fecesHumansSouth Korea
NCC 533GCF_000008065.134.61.9918210HumansHumansNA
DPC 6026GCF_000204985.134.81.9719000Small intestineSwineNA
N6.2GCF_000498675.134.21.8917180RatRatNA
FI9785GCF_000091405.134.51.7617652PoultryPoultryUK
pf01GCA_000219475.334.61.9318462Piglet fecesSwineKorea
Note: the genomes of strains N5 and N7 were determined in this study. ‘NA’ denotes no exact geographic information was obtained based on the information listed in the NCBI Biosample database.
Table 2. Carbohydrate fermentation profile of L. Johnsonii N5 and N7 as assessed by API 50 CH test strips.
Table 2. Carbohydrate fermentation profile of L. Johnsonii N5 and N7 as assessed by API 50 CH test strips.
SubstrateResultSubstrateResult
N5N7N5N7
ControlEsculine
GlycerolSalicine++
ErythritolD-cellobiose++
D-arabinoseD-maltose++
L-arabinoseD-lactose
D-riboseD-melibiose++
D-xyloseD-saccharose++
L-xyloseD-threalose
D-adonitolInulin
Methyl-β-dxylopyranoside++D-melezitose++
D-galactose++D-raffinose
D-glucose++Starch
D-fructoseGlycogene
D-mannoseXylitol++
L-sorboseGentiobiose
L-rhamnoseD-turanose
DulcitolD-lyxose
InositolD-tagatose
D-mannitolD-fucose
D-sorbitolL-fucose
Methyl-αD-mannopyranosideD-arabitol
Methyl-αD-glucopyranoside++L-arabitol
N-acetylglucosaminePotassium gluconate
AmygdalinePotassium 2-cetogluconate
Arbutine++Potassium 5-cetogluconate
Notes: “−” denotes the substrate cannot be utilized, and “+” denotes the substrate can be utilized.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, K.; Wang, Y.; Gu, L.; Yu, J.; Liu, Q.; Zhang, R.; Liang, G.; Chen, H.; Gu, F.; Liu, H.; et al. Characterization of Probiotic Properties and Whole-Genome Analysis of Lactobacillus johnsonii N5 and N7 Isolated from Swine. Microorganisms 2024, 12, 672. https://doi.org/10.3390/microorganisms12040672

AMA Style

Wang K, Wang Y, Gu L, Yu J, Liu Q, Zhang R, Liang G, Chen H, Gu F, Liu H, et al. Characterization of Probiotic Properties and Whole-Genome Analysis of Lactobacillus johnsonii N5 and N7 Isolated from Swine. Microorganisms. 2024; 12(4):672. https://doi.org/10.3390/microorganisms12040672

Chicago/Turabian Style

Wang, Kun, Yu Wang, Lifang Gu, Jinyan Yu, Qianwen Liu, Ruiqi Zhang, Guixin Liang, Huan Chen, Fang Gu, Haoyu Liu, and et al. 2024. "Characterization of Probiotic Properties and Whole-Genome Analysis of Lactobacillus johnsonii N5 and N7 Isolated from Swine" Microorganisms 12, no. 4: 672. https://doi.org/10.3390/microorganisms12040672

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

Article Metrics

Back to TopTop