Skip to main content

Insights about the epidemiology of Salmonella Typhimurium isolates from different sources in Brazil using comparative genomics

Abstract

Background

Salmonella enterica subsp. enterica serovar Typhimurium (S. Typhimurium) is an important zoonotic agent worldwide. The aim of this work was to compare genetically 117 S. Typhimurium isolated from different sources over 30 years in Brazil using different genomics strategies.

Results

The majority of the 117 S. Typhimurium strains studied were grouped into a single cluster ( 90%) by the core genome multilocus sequence typing and ( 77%) by single copy marker genes. The phylogenetic analysis based on single nucleotide polymorphism (SNP) grouped most strains from humans into a single cluster ( 93%), while the strains isolated from food and swine were alocated into three clusters. The different orthologous protein clusters found for some S. Typhimurium isolated from humans and food are involved in metabolic and regulatory processes. For 26 isolates from swine the sequence types (ST) 19 and ST1921 were the most prevalent ones, and the ST14, ST64, ST516 and ST639 were also detected. Previous results typed the 91 S. Typhimurium isolates from humans and foods as ST19, ST313, ST1921, ST3343 and ST1649. The main prophages detected were: Gifsy-2 in 79 (67.5%) and Gifsy-1 in 63 (54%) strains. All of the S. Typhimurium isolates contained the acrA, acrB, macA, macB, mdtK, emrA, emrB, emrR and tolC efflux pump genes.

Conclusions

The phylogenetic trees grouped the majority of the S. Typhimurium isolates from humans into a single cluster suggesting that there is one prevalent subtype in Brazil. Regarding strains isolated from food and swine, the SNPs’ results suggested the circulation of more than one subtype over 30 years in this country. The orthologous protein clusters analysis revealed unique genes in the strains studied mainly related to bacterial metabolism. S. Typhimurium strains from swine showed greater diversity of STs and prophages in comparison to strains isolated from humans and foods. The pathogenic potential of S. Typhimurium strains was corroborated by the presence of exclusive prophages of this serovar involved in its virulence. The high number of resistance genes related to efflux pumps is worrying and may lead to therapeutic failures when clinical treatment is needed.

Background

Nontyphoidal Salmonella (NTS) strains have been an important enteric agent transmitted mainly by contaminated foods worldwide [1]. According to Kirk and collaborators [2], it was estimated that 153 million infections and 56,969 deaths occurred around the globe due to salmonellosis in 2010. Moreover, data from the Centers for Disease Control and Prevention (CDC), estimated that 1.35 million infections, 26,500 hospitalizations and 420 deaths occur in the United States every year due to Salmonella [3].

In Brazil, Salmonella has been the first or second most common foodborne pathogen isolated from outbreaks in recent years [4]. However, until now there are few published studies that have characterized the possible differences between Brazilian Salmonella enterica subsp. enterica serovar Typhimurium (S. Typhimurium) strains isolated from human, food and animal sources by whole genome sequencing (WGS).

Salmonella Typhimurium is one of the main Salmonella generalist serovar, which has been isolated from pork in Europe, Oceania, Asia and North America, from poultry in North America and Oceania, from beef in Africa, Latin America and Europe, and from seafood in Europe [5]. Therefore, this serovar has been transmitted from animals and humans in different parts of the world and is characterized as a zoonotic agent causing losses of million of dollars for the pork, poultry and beef producing industry [1, 6].

According to the CDC, S. Typhimurium can also infect domestic pets and recently was responsible for an outbreak linked to contact with small pet turtles that affected 35 people from nine states and generated 11 hospitalizations [7].

WGS has been more accessible in the last few years and is used for molecular characterization studies [8]. Furthermore, different phylogenetic strategies can be performed after sequencing, such as construction of phylogenetic trees based on the core genome multilocus sequencing typing (cgMLST), from single copy marker genes and from single nucleotide polymorphism (SNPs), besides comparison and analysis of orthologous protein clusters (OrthoVenn) and verification of the sequence type (ST) through multilocus sequence typing (MLST) [9,10,11]. In addition, it has been possible to characterize the different prophages that contribute to Salmonella pathogenicity including identification of genes known to have functions such as virulence, metabolism and signaling [12].

It is important to emphasize that the monitoring of resistant NTS strains has been of great importance due to its continued emergence worldwide [13, 14]. According to Jajere, multidrug resistant (MDR) Salmonella has been a serious public health problem because it may lead to treatment failure when the uses of antimicrobial drugs are necessary [14]. In the United States, it was estimated that 212,500 infections and 70 deaths occur due to drug resistant NTS every year [13].

It is known that hundreds of genes can confer resistance to antibiotics in NTS and some were previously described for the S. Typhimurium strains isolated from humans and different foods in Brazil including genes related to resistance to aminoglycosides, tetracyclines, sulfonamides, trimethoprim, beta lactams, fluoroquinolones, phenicol and macrolides [15]. However, antibiotic resistance is multifactorial and little is known about resistance genes related to efflux pumps, which can be an important factor that confers resistance to some antibiotics, such as fluoroquinolones, beta lactams, macrolides and aminoglycosides [15, 16].

The aim of this work was to compare genetically S. Typhimurium isolates from humans, food and swine in Brazil from over 30 years using different genomics strategies, such as phylogenetic trees, protein orthologous clusters analysis, MLST, prophages and resistance genes related to efflux pump.

Results

cgMLST

The cgMLST grouped the 120 S. Typhimurium genomes studied, which included the three references analysed in two main groups designated A and B (Fig. 1). Cluster A comprised 12 genomes of ST19 isolated from humans. Cluster B comprised a total of 108 genomes comprising strains isolated from humans, different foods and swine of ST19, ST1649, ST3343, ST1921 and ST313 in the case of strains isolated from humans and food, besides ST19, ST639, ST14, ST516, ST64 and ST1921 concerning strains isolated from swine. All three references were allocated in Cluster B. The CFSAN033848 and CFSAN033855 genomes isolated from humans were genetically distinct and did not group closely to any other isolates.

Fig. 1
figure 1

Phylogenetic analysis with cgMLST profiles based on soft core of 3002 genes selected for 117 Salmonella Typhimurium genomes isolated from humans (n = 43), foods (n = 48) and swine (n = 26) in Brazil

Phylogenetic tree (ggTree) and orthologous protein clusters analysis

The ggTree grouped the 120 S. Typhimurium genomes studied, which included the three references analysed, in three groups designated A, B and C with cluster A subdivided in A.1 and A.2, cluster B subdivided in B.1 and B.2 (Fig. 2). Cluster A.1 comprised 84 genomes of ST19, ST1649, ST14, ST516, ST639, ST64, ST313, ST3343 and ST1921 isolated from humans, diverse foods and swine and the reference genomes. Cluster A.2 comprised nine genomes of ST19 isolated from humans, food and swine. Cluster B.1 comprised 20 genomes of ST19 from food and swine. Cluster B.2 comprised four genomes of ST19 isolated from human and food. Cluster C comprised three genomes of ST19 isolated from food and swine.

Fig. 2
figure 2

Phylogenetic analysis based on a list of single copy marker genes by ggTree for 117 Salmonella Typhimurium genomes isolated from humans (n = 43), foods (n = 48) and swine (n = 26) in Brazil

The orthologous protein clusters analysis was performed for the genomes that were more related to LT2, 14028S and D23580 references (Fig. 2). The comparisons indicated the orthologous protein clusters presented in the genomes of the strains of this study and absent in the references. The different unique orthologous protein clusters found are involved in metabolic and regulatory processes showed in detail in Table 1.

Table 1 Unique orthologous protein clusters in some selected S. Typhimurium strains in comparison to reference genomes

snpTree

The snpTree grouped the 120 S. Typhimurium genomes studied, which included the three references analysed, in three groups designated A, B and C (Fig. 3). Cluster A comprised 81 genomes of ST19, ST14, ST516, ST639, ST64, ST1649, ST313, ST3343 and ST1921 isolated from humans, food and swine, plus all three references. Cluster B comprised 28 genomes including one strain isolated from human and 27 strains isolated from different foods and swine of ST19. Cluster C comprised seven genomes including one strain isolated from human and six strains isolated from food of ST19. The CFSAN033890 genome isolated from human was genetically distinct and did not group closely to any other isolates.

Fig. 3
figure 3

Phylogenetic analysis based on SNPs from whole genome sequencing by snpTree for 117 Salmonella Typhimurium genomes isolated from humans (n = 43), foods (n = 48) and swine (n = 26) in Brazil

MLST

Of the 26 S. Typhimurium strains isolated from swine studied, 16 (61.5%) belonged to the ST19, three (11.5%) to the ST1921, two (7.6%) to the ST14, two (7.6%) to the ST64, one (3.8%) to the ST516, one (3.8%) to the ST639 and one isolate did not match any known ST type. The 91 S. Typhimurium isolates from humans and foods were previously typed and the ST19 (84.6%) was the most prevalent, followed by the ST313 (9.9%), ST1921 (2.2%), ST3343 (2.2%) and ST1649 (1.1%) as described in Almeida et al. [19].

Prophages detection

The Gifsy-2 prophage was detected in 79 (67.5%) S. Typhimurium isolates, Gifsy-1 in 63 (54%), Salmon 118970_sal3 in 46 (39%) and Haemop—HP1 in 21 (18%). Two dozen other prophages were also detected in the genomes studied and are described in detail in Table 2.

Table 2 Proportion of intact prophages detected in the 117 Salmonella Typhimurium studied isolated in Brazil

Efflux pumps

The acrA, acrB, macA, macB, mdtK, emrA, emrB, emrR and tolC genes were detected in the 117 (100%) S. Typhimurium studied. The mdsA and mdsB genes were detected in 91 (100%) S. Typhimurium isolates from humans and different foods, but in only 18 (69.2%) S. Typhimurium isolates from swine. The mdfA gene was detected in 26 (100%) isolates from swine, 39 (81.2%) isolates from food and 18 (42%) isolates from humans. Finally, the cmlA1 gene was detected only in isolates from swine 05 (19.2%). The percentage of query cover and identity for all genes ranged between 72 and 100 and 87–100, respectively (Table 3).

Table 3 Frequencies of resistance genes related to efflux pumps in the 117 Salmonella Typhimurium studied

Discussion

In this study, 117 S. Typhimurium isolates from humans (n = 43), food (n = 48) and swine (n = 26) in Brazil were compared using genomic analyses, such as phylogenetic trees, orthologous protein clusters detection, MLST analysis, and blast identification of prophages and resistance genes related to efflux pumps.

The majority of the 117 S. Typhimurium strains studied were grouped into a single cluster ( 90%) by the core genome multilocus sequence typing and ( 77%) by single copy marker genes (Figs. 1 and 2). The phylogenetic analysis based on single nucleotide polymorphism (SNPs) grouped most strains from humans into a single cluster ( 93%), while the strains isolated from food and swine were grouped into three clusters (Fig. 3). Therefore, snpTree was more efficient at discriminating S. Typhimurium isolates from swine and different foods in Brazil.

It is important to mention that the present study provided additional information about S. Typhimurium strains isolated from humans, food and swine in Brazil because such strains have rarely been studied in a one health perspective combining all available data [15, 17,18,19].

Previous studies performed by our research group using different molecular typing techniques (PFGE, MLVA and CRISPR-MVLST) and a SNP-based tree by the CFSAN pipeline corroborated with the finding of snpTree by CSI Phylogeny 1.4. indicating the possible presence of a prevalent subtype for S. Typhimurium strains isolated from humans and with more than one circulating subtype for strains isolated from food [15, 17,18,19].

According to Jensen, homologous genes can be divided into orthologous and paralogs genes [20]. Orthologous genes originated from a common ancestor during speciation events and keep the same function, while, paralogs genes originated from duplication events and do not maintain the same function [21].

Therefore, the OrthoVenn2 is a web server capable to annotate and compare orthologous protein clusters from the whole genome among different species [21]. In the present study, S. Typhimurium genomes were compared to LT2, 14028S and D23580 references and had their unique protein orthologous clusters determined (Fig. 2). All S. Typhimurium isolates compared to LT2 (Comparisons 1, 2, 3 and 4) were composed of ST19 isolated from humans in the São Paulo State before the 1990s. There were some unique orthologous protein clusters, including transposition (DNA-mediated), transposition, viral genome integration into host DNA and Trehalose transport that were commonly present in these strains, but absent in the corresponding LT2 reference strain. The S. Typhimurium isolates compared to 14028S contained ST19 strains isolated from food in the Rio Grande do Sul, Santa Catarina and Bahia States between 2006 and 2012 (Comparison 5). The S. Typhimurium isolates compared to D23580 contained ST313 and ST19 strains isolated from humans and food in the São Paulo and Paraná States between 1995 and 2010 (Comparisons 6 and 7).

The different orthologous protein clusters found are involved in metabolic and regulatory processes, such as transposition, DNA replication, cell adhesion, formate oxidation, trehalose transport, lyase activity and response to mercury ion. These results showed that despite being of the same serovar there are unique orthologous protein clusters in the strains studied in comparison to the reference strains which were maintained in these S. Typhimurium strains during natural selection and adaptation (Table 1).

In this study, MLST was performed only for swine isolates, because the STs for humans and food isolates were previously described by Almeida et al. [22]. Of the 26 S. Typhimurium strains isolated from swine studied, 16 (61.5%) belonged to the ST19, three (11.5%) to the ST1921, two (7.6%) to the ST14, two (7.6%) to the ST64, one (3.8%) to the ST516, one (3.8%) to the ST639 and one did not have its ST detected.

Previous works showed that the ST19 was the most common ST found for strains of human and food origins, with ST313 being the second most prevalent and ST1921, ST3343 and ST1649 were also detected among these strains [22]. S. Typhimurium isolates from swine showed greater diversity in the seven housekeeping genes studied despite having a lower number of strains (n = 26) in comparison to the number of S. Typhimurium strains isolated from humans (n = 43) and food (n = 48). ST19 was the most commonly observed in swine, with ST1921 the second most prevalent and with ST14, ST64, ST516 and ST639 also observed.

For ST19 it has been reported 29,572 Salmonella isolated from human, reptile, ovine, swine, poultry, food and bovine from France, Mexico, China, Germany, Scotland, Portugal, Qatar, Korea, Ireland, United States (US), United Kingdom (UK) and Denmark according to the Enterobase (12/15/2020). The ST313 has been linked to 3049 samples isolated predominantly from humans in Kenya, Ethiopia, Zimbabwe, Malawi, Mali and Nigeria [10].

Moreover, ST1649, ST3343 and ST1921 were found for 16, 4 and 7 strains respectively, isolated from humans, livestocks, food and swine in Venezuela, Ireland, US, UK, Colombia, Ecuador, Vietnam and Brazil [10]. Finally, ST516, ST64, ST639 and ST14 were linked to 370, 3850, 237 and 2149 strains respectively, isolated from humans, poultry, food, aquatic animals, reptiles and the environment in the US, Mexico, Senegal, Germany, Portugal, Qatar, Canada, UK, India, Ghana, Thailand, Malaysia, Malta, Vietnam, Pakistan, Greece, France, Germany, China, Denmark, Scotland, Norway and South Korea [10].

It is important to emphasize that the classic MLST sequencing scheme uses only seven housekeeping genes to determine a sequence type (ST) from the nucleotides differences found in the sequences of all alleles [23]. Furthermore, cgMLST focuses on the nucleotide differences between the set of 3002 conserved genes of Salmonella genus [10]. It is known that the ST19 has been more prevalent in S. Typhimurium strains which causes predominantly gastroenteritis worldwide, suggesting that in the tree based on cgMLST there is a greater diversity in the 3002 conserved genes because S. Typhimurium strains isolated from humans of this ST were found in the cluster A and B (Fig. 1).

In the present study, the Gifsy-2 prophage was detected in 79 (67.5%) S. Typhimurium isolates, Gifsy-1 in 63 (54%), Salmon 118970_sal3 in 46 (39%) and Haemop—HP1 in 21 (18%). Specifically, Gifsy-1 prophage was detected in 10 (23%) S. Typhimurium strains isolated from humans, 34 (71%) strains isolated from different foods and in 19 (73%) strains isolated from swine (Table 2). Gifsy-2 prophage was detected in 37 (86%) S. Typhimurium strains isolated from humans, 33 (68.7%) strains isolated from foods and 9 (34.6%) strains isolated from swine (Table 2).

It is important to be mentioned that Gifsy prophages carry genes that are related to virulence of S. Typhimurium in the host [24, 25]. The Gifsy-1 prophage encodes three genes involved in the intracellular survival of Salmonella spp. in the host, denominated gogB (leucine-rich repeat protein), sarA (anti-inflammatory response activator) and pagK2. In the same way, the Gifsy-2 prophage encodes a superoxide dismutase (sodC1) that contributes to the survival of Salmonella spp. destroying the toxic radicals of the host macrophages [12, 26]. It is important to emphasize that Gifsy prophages have been found only in S. Typhimurium strains, as well as the Fels-1 and Fels-2 prophages [25]. In the present study, Fels prophages were detected in four S. Typhimurium strains isolated from humans, two strains isolated from food and four strains isolated from swine (Table 2).

According to Brussow et al. [27], the Fels-1 prophage encodes the sodCIII and nanH genes related to the production of superoxide dismutase and neuraminidase in S. Typhimurium, respectively. Furthermore, the Fels-2 prophage carries genes that are apparently related to regulation and adhesion of S. Typhimurium to host cells [12].

The Gifsy and Fels prophages have already been described in S. Typhimurium isolated in various parts of the world, such as Australia, Europe, China, among others [28,29,30]. It is important to emphasize that other prophages were also found in the S. Typhimurium strains studied including Salmon 118970_sal3 and Haemop—HP1 (Table 2). Moreover, two dozen other prophages were detected in the S. Typhimurium strains studied, but there is less information about them related to pathogenicity and/or virulence of this serovar (Table 2).

In addition, S. Typhimurium isolates from swine showed 6 (23.1%) unique prophages despite having a lower number of strains analysed (n = 26) in comparison to S. Typhimurium strains isolated humans (n = 43) and food (n = 48) that presented 7 (16.3%) and 3 (6.25%) unique prophages, respectively, suggesting the greater diversity in these mobile genetic element for S. Typhimurium strains isolated from swine in Brazil (Table 2).

Resistance to multiple drugs in bacteria has been a serious public health problem worldwide [31]. It is known that there are four main mechanisms that can cause this resistance, such as target alteration, drug inactivation, decreased permeability and drug expulsion through the production of efflux pumps [32].

In the present study, the acrA, acrB, macA, macB, mdtK, emrA, emrB, emrR, tolC, mdsA, mdsB, mdfA and cmlA1 genes were detected among the S. Typhimurium strains isolated from humans, food and swine. All of the isolates contained the acrA, acrB, macA, macB, mdtK, emrA, emrB, emrR and tolC genes (Table 3). Other genes related to production efflux pump, such as oqxAB and floR were previously reported in [15].

The AcrAB efflux system has been described as responsible for the intrinsic resistance to many antibiotics that can be used in medical practice for the treatment of S. Typhimurium, such as fluoroquinolones and beta-lactams [16]. According to the World Health Organization (WHO), Salmonella spp. was described as a high priority category pathogen in fluoroquinolones resistance of the Global Priority Pathogens List [31].

The macA and macB genes encode proteins that characterize an efflux pump related to macrolides resistance [33, 34]. According to the Universal Protein Resource (UniProt), the mdtK, emrA, emrB, emrR, mdsA, mdsB, mdfA and cmlA1 genes encode mainly proteins involved in multidrug efflux transporter and confers resistance to different antibiotics, such as aminoglycosides, tetracyclines, novobiocin, nalidixic acid, chloramphenicol and norfloxacin [34, 35]. Furthermore, the tolC gene has been described as important for the formation of some multidrug efflux systems (AcrAB, MacAB, EmrAB and MdsAB) in S. Typhimurium [35].

Conclusions

The phylogenetic trees grouped the majority of the S. Typhimurium isolates from humans into a single cluster suggesting that there is one prevalent subtype in Brazil. Regarding strains isolated from food and swine, the results by SNPs analysis suggested the circulation of more than one subtype over 30 years in this country. The orthologous protein clusters analysis revealed unique genes in the strains studied mainly related to bacterial metabolism. S. Typhimurium isolates from swine showed greater diversity of STs and prophages in comparison to S. Typhimurium strains isolated from humans and food. The pathogenic potential of S. Typhimurium strains was corroborated by the presence of exclusive prophages of this serovar involved in their virulence. The high number of resistance genes related to efflux pump is worrying and may cause therapeutic failures when clinical treatment is needed. Altogether, this study provided relevant data on the genomic characterization of S. Typhimurium strains isolated from different sources in Brazil using WGS.

Methods

Bacterial strains

A total of 117 S. Typhimurium strains isolated from humans (43), food (48) and swine (26) between 1983 and 2013 in Brazil were studied (Table 4). These strains were selected from the collections of the Adolfo Lutz Institute of Ribeirão Preto (IAL-RP), of the Oswaldo Cruz Foundation from Rio de Janeiro (FIOCRUZ-RJ) and of the Brazilian Agricultural Research Corporation (EMBRAPA).

Table 4 Characteristics of the 117 Salmonella Typhimurium strains studied isolated from different sources in Brazil

Whole genome sequencing

The DNA of the 117 S. Typhimurium strains was extracted according to Campioni and Falcão using phenol-chloroform-isoamyl alcohol method [36]. Libraries were prepared using 1 ng of genomic DNA with the Nextera XT DNA library preparation kit (Illumina, San Diego, CA, USA) and the genomes were sequenced using the NextSeq 500 desktop sequencer with the NextSeq 500/500 high-output version 2 kit (Illumina) for 2 × 151 cycles according to the manufacturer’s instructions at the U.S. Food and Drug Administration (FDA), College Park, Maryland, USA. The genomes were assembled using the software SPAdes and CLC Genomics Workbench version 10.0.1 [37] and the quality of the assemblies were evaluated using the software QUAST [38]. The genomes ranged from 4.6 to 5.1 Mb in size, as described for other Salmonella strains [39]. Sequencing generated an average G+C content of 52.04%, which is similar to that reported previously for other Salmonella isolates [40]. The number of contigs per assembly for each isolate ranged between 47 and 827. Finally, the coverage (×) ranged from 13× to 753×. Detailed information on the sequencing of the 117 S. Typhimurium genomes can be found in Almeida et al. and Seribelli et al. [41, 42].

cgMLST

The cgMLSTFinder 1.1 analysis was determined from a set of reads for all 117 S. Typhimurium genomes and three different references of this serovar were chosen, which included LT2, 14028S and D23580 and compared using the services of the center for genomic epidemiology for Salmonella (Enterobase) available at https://cge.cbs.dtu.dk/services/cgMLSTFinder/ [10].

Phylogenetic tree (ggTree) and orthologous protein clusters analysis

Three different references of S. Typhimurium serovar were chosen, which included LT2 (GCF_000006945), 14028S (GCF_000022165) and D23580 (GCF_900538085), all with fully closed deposited genomes. To evaluate the evolutionary distance between the sequenced genomes and the three reference strains, a neighbor-joining tree was built with the ezTree algorithm [11] and ggTree R package [43, 44] (Fig. 2). The ezTree has been described as an automated pipeline based in the single copy marker genes identification to construct a phylogenetic tree for a set of input genomes [11]. Additional characterization of the orthologous protein clusters for some of the key S. Typhimurium strains were performed. The phylogroups selected included: Comparison 1—LT2 with 11 genomes (CFSAN033873, CFSAN033871, CFSAN033874, CFSAN033859, CFSAN033869, CFSAN033872, CFSAN033863, CFSAN033854, CFSAN033867, CFSAN033866 and CFSAN033868); Comparison 2—LT2 with two genomes (CFSAN033855 and CFSAN033848); Comparison 3—LT2 with six genomes (CFSAN033856, CFSAN033860, CFSAN033861, CFSAN033857, CFSAN033858 and CFSAN033851); Comparison 4—LT2 with three genomes (CFSAN033852, CFSAN033849 and CFSAN033850); Comparison 5—14028S with seven genomes (CFSAN033930, CFSAN033931, CFSAN033919, CFSAN033924, CFSAN033922, CFSAN033929 and CFSAN033909); Comparison 6—D23580 with nine genomes (CFSAN033882, CFSAN033877, CFSAN033887, CFSAN033886, CFSAN033876, CFSAN033881, CFSAN033894, CFSAN033891 and CFSAN033884); Comparison 7—D23580 with 11 genomes (CFSAN033882, CFSAN033877, CFSAN033887, CFSAN033886, CFSAN033876, CFSAN033881, CFSAN033894, CFSAN033891, CFSAN033884, CFSAN033921 and CFSAN033883) via OrthoVenn2 [21] in order to determine unique features and metabolic pathways defining each group.

SNP tree

The phylogenetic tree based on SNPs of the whole genome sequencing was performed by CSI Phylogeny 1.4 (Call SNPs & Infer Phylogeny) of the Center for Genomic Epidemiology at https://cge.cbs.dtu.dk/services/CSIPhylogeny/—following the parameters: select min. depth at SNP positions 10×, select min. relative depth at SNP positions 10%, select minimum distance between SNPs (prune) 10 bp, select min. SNP quality 30, select min. read mapping quality 25 and select min. Z-score 1.96 [45]. The SNPs matrix included was a maximum of 30,873 SNPs among all S. Typhimurium strains studied.

Multilocus sequence typing (MLST)

MLST was performed in the present study for the 26 S. Typhimurium isolates from swine using the MLST 2.0 of the Center for Genomic Epidemiology for Salmonella enterica available in https://cge.cbs.dtu.dk/services/MLST/ [46]. The seven housekeeping genes included: aroC, dnaN, hemD, hisD, purE, sucA and thrA [23, 46]. The STs of the S. Typhimurium isolates from humans and different foods were previously described in Almeida et al. [22] and were performed in the same way as described above.

Prophages detection

The genomes of all 117 S. Typhimurium strains were used to search the prophages by PHAge Search Tool Enhanced Release (PHASTER) that is an online platform for the rapid identification and annotation of prophages sequences in bacterial genomes and plasmids available in http://phaster.ca/ [47].

Efflux pumps

The genomes of all 117 S. Typhimurium strains were used to search for resistance genes related to efflux pump. Resistance gene identifier (RGI) is part of the Comprehensive Antibiotic Resistance Database (CARD) and was performed with high quality/coverage (includes contigs > 20,000 bp and excludes prediction of partial genes). Software is available at https://card.mcmaster.ca/analyze/rgi [48].

Availability of data and materials

The data from 117 S. Typhimurium genomes were deposited in the GenBank (NCBI) under the identification numbers: LVHC00000000, LVHB00000000, LVHA00000000, LVGZ00000000, LVGY00000000, LVGX00000000, LVGW00000000, MABI00000000, LVGV00000000, LVGU00000000, LVGT00000000, LUJG00000000, LVGS00000000, LVGR00000000, LVGQ00000000, LVGP00000000, LVGO00000000, LVGN00000000, LVGM00000000, LVGL00000000, LUJF00000000, LVGK00000000, LVGJ00000000, LVGI00000000, LVGH00000000, LVGG00000000, LVGF00000000, LUJE00000000, LVGE00000000, LVGD00000000, LUJD00000000, LVGC00000000, LVGB00000000, LVGA00000000, LVFZ00000000, LVFY00000000, LVFX00000000, LUJC00000000, LUJB00000000, LUJA00000000, LVFW00000000, LUIZ00000000, LVFV00000000, LVFU00000000, LUIY00000000, LVFT00000000, LUIX00000000, LUIW00000000, LVFS00000000, LVFR00000000, LUIV00000000, LUIU00000000, LUIT00000000, LVFQ00000000, LUIS00000000, LUIR00000000, LUIQ00000000, LUIP00000000, LUIO00000000, LVFP00000000, LUIN00000000, LUIM00000000, LUIL00000000, LUIK00000000, LVFO00000000, LUIJ00000000, LUII00000000, LUIH00000000, LVFN00000000, LUIG00000000, LUIF00000000, LUIE00000000, LVFM00000000, LUID00000000, LUIC00000000, LVFL00000000, LVFK00000000, LUIB00000000, LUIA00000000, LUHZ00000000, LVFJ00000000, LUHY00000000, LUHX00000000, LVFI00000000, LUHW00000000, LUHV00000000, LUHU00000000, LUHT00000000, LUHS00000000, LUHR00000000, LVFH00000000, SRR8291813, SRR8291805, SRR8291802, SRR8291817, SRR8291806, SRR8291814, PHJE00000000, PHJD00000000, PHJC00000000, PIJC00000000, PHJB00000000, PHJA00000000, PHIZ00000000, PHIY00000000, PHIX00000000, PHIW00000000, PHIV00000000, PHIU00000000, PHIT00000000, PHIS00000000, PHIR00000000, PHIQ00000000, PHIP00000000, PHIO00000000, PHIN00000000, PHIM00000000, released under the project PRJNA186035 (https://www.ncbi.nlm.nih.gov/bioproject/186035).

Abbreviations

NTS :

Nontyphoidal Salmonella

CDC :

Centers for Disease Control and Prevention

WGS :

Whole genome sequencing

cgMLST :

Core genome multilocus sequence typing

SNP :

Single nucleotide polymorphism

ST :

Sequence type

MLST :

Multilocus sequence typing

MDR :

Multidrug resistant

DNA :

Deoxyribonucleic acid

WHO :

World Health Organization

UniProt :

Universal Protein Resource

FDA :

Food and Drug Administration

PHASTER :

PHAge Search Tool Enhanced Release

RGI :

Resistance gene identifier

CARD :

Comprehensive Antibiotic Resistance Database

IAL-RP :

Adolfo Lutz Institute of Ribeirão Preto

FIOCRUZ-RJ :

Oswaldo Cruz Foundation from Rio de Janeiro

EMBRAPA :

Brazilian Agricultural Research Corporation

References

  1. World Health Organization (WHO).  Salmonella (non-typhoidal) https://www.who.int/news-room/fact-sheets/detail/salmonella-(non-typhoidal). Accessed 24 Nov 2020.

  2. Kirk MD, Pires SM, Black RE, Caipo M, Crump JA, Devleesschauwer B, et al. World Health Organization estimates of the global and regional disease burden of 22 foodborne bacterial, protozoal, and viral diseases, 2010: a data synthesis. PLoS Med. 2015;12(12):e1001921. https://doi.org/10.1371/journal.pmed.1001921.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Centers for Disease Control and Prevention (CDC). Salmonella. https://www.cdc.gov/salmonella/index.html. Accessed 24 Nov 2020.

  4. Surtos de Doenças. Transmitidas por Alimentos no Brasil Informe. 2018.  https://portalarquivos2.saude.gov.br/images/pdf/2019/fevereiro/15/Apresenta----o-Surtos-DTA---Fevereiro-2019.pdf. Accessed 24 Nov 2020.

  5. Ferrari RG, Rosario DKA, Cunha-Neto A, Mano SB, Figueiredo EES, Conte-Junior CA. Worldwide epidemiology of Salmonella serovars in animal-based foods: a meta-analysis. Appl Environ Microbiol. 2019. https://doi.org/10.1128/AEM.00591-19.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Manual Técnico de. Diagnóstico Laboratorial da Salmonella spp. Ministério da Saúde. https://portalarquivos2.saude.gov.br/images/pdf/2014/dezembro/15/manual-diagnostico-salmonella-spp-web.pdf. Accessed 24 Nov 2020.

  7. Outbreak of Salmonella infections linked to small pet turtles. https://www.cdc.gov/salmonella/typhimurium-1-20/index.html. Accessed 24 Nov 2020.

  8. Shivani C, Abha K, Alka G, Sampat N. Comparative genome analysis of three pathogenic strains of E. coli, Salmonella and Shigella. IJSRR. 2015;4(1):68–80.

    CAS  Google Scholar 

  9. Wang Y, Coleman-Derr D, Chen G, Gu YQ. OrthoVenn: a web server for genome wide comparison and annotation of orthologous clusters across multiple species. Nucleic Acids Res. 2015;43(W1):W78–84. https://doi.org/10.1093/nar/gkv487.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Alikhan NF, Zhou Z, Sergeant MJ, Achtman M. A genomic overview of the population structure of Salmonella. PLoS Genet. 2018;14(4):e1007261. doi:https://doi.org/10.1371/journal.pgen.1007261.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wu YW. ezTree: an automated pipeline for identifying phylogenetic marker genes and inferring evolutionary relationships among uncultivated prokaryotic draft genomes. BMC Genom. 2018;19(Suppl 1):921. doi:https://doi.org/10.1186/s12864-017-4327-9.

    Article  CAS  Google Scholar 

  12. Wahl A, Battesti A, Ansaldi M. Prophages in Salmonella enterica: a driving force in reshaping the genome and physiology of their bacterial host? Mol Microbiol. 2019;111(2):303–16. https://doi.org/10.1111/mmi.14167.

    Article  CAS  PubMed  Google Scholar 

  13. Antibiotic resistance threats report. Drug-resistant nontyphoidal Salmonella. https://www.cdc.gov/drugresistance/pdf/threats-report/nt-salmonella-508.pdf. Accessed 24 Nov 2020.

  14. Jajere SM. A review of Salmonella enterica with particular focus on the pathogenicity and virulence factors, host specificity and antimicrobial resistance including multidrug resistance. Vet World. 2019;12(4):504–21. https://doi.org/10.14202/vetworld.2019.504-521.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Almeida F, Seribelli AA, Medeiros MIC, Rodrigues DDP, de MelloVarani A, Luo Y, et al. Phylogenetic and antimicrobial resistance gene analysis of Salmonella Typhimurium strains isolated in Brazil by whole genome sequencing. PLoS ONE. 2018;13(8):e0201882. https://doi.org/10.1371/journal.pone.0201882.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Piddock LJV. Understanding the basis of antibiotic resistance: a platform for drug discovery. Microbiology. 2014;160(Pt11):2366–73. https://doi.org/10.1099/mic.0.082412-0.

    Article  CAS  PubMed  Google Scholar 

  17. Almeida F, Medeiros MIC, Rodrigues DDP, Falcão JP. Genotypic diversity, pathogenic potential and the resistance profile of Salmonella Typhimurium strains isolated from humans and food from 1983 to 2013 in Brazil. J Med Microbiol. 2015;64(11):1395–407. https://doi.org/10.1099/jmm.0.000158.

    Article  CAS  PubMed  Google Scholar 

  18. Almeida F, Medeiros MIC, Kich JD, Falcão JP. Virulence-associated genes, antimicrobial resistance and molecular typing of Salmonella Typhimurium strains isolated from swine from 2000 to 2012 in Brazil. J Appl Microbiol. 2016;120(6):1677–90. https://doi.org/10.1111/jam.13110.

    Article  CAS  PubMed  Google Scholar 

  19. Almeida F, Medeiros MIC, Rodrigues DDP, Allard MW, Falcão JP. Molecular characterization of Salmonella Typhimurium isolated in Brazil by CRISPR-MVLST. J Microbiol Methods. 2017;133:55–61. https://doi.org/10.1016/j.mimet.2016.12.020.

    Article  CAS  PubMed  Google Scholar 

  20. Jensen RA. Orthologs and paralogs—we need to get it right. Genome Biol. 2001;2(8):1–3. https://doi.org/10.1186/gb-2001-2-8-interactions1002.

    Article  Google Scholar 

  21. Xu L, Dong Z, Fang L, Luo Y, Wei Z, Guo H, et al. OrthoVenn2: a web server for whole-genome comparison and annotation of orthologous clusters across multiple species. Nucleic Acids Res. 2019;47(W1):W52–8. https://doi.org/10.1093/nar/gkz333.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Almeida F, Seribelli AA, da Silva P, Medeiros MIC, Dos Prazeres Rodrigues D, Moreira CG, et al. Multilocus sequence typing of Salmonella Typhimurium reveals the presence of the highly invasive ST313 in Brazil. Infect Genet Evol. 2017;51:41–4. https://doi.org/10.1016/j.meegid.2017.03.009.

    Article  CAS  PubMed  Google Scholar 

  23. Achtman M, Wain J, Weill FX, Nair S, Zhou Z, Sangal V, et al. Multilocus sequence typing as a replacement for serotyping in Salmonella enterica. PloS Pathog. 2012;8(6):e1002776. doi:https://doi.org/10.1371/journal.ppat.1002776.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Klumpp J, Fuchs TM. Identification of novel genes in genomic islands that contribute to Salmonella Typhimurium replication in macrophages. Microbiology. 2007;153(Pt 4):1207–20. https://doi.org/10.1099/mic.0.2006/004747-0.

    Article  CAS  PubMed  Google Scholar 

  25. Ngoi ST, Yap KP, Thong KL. Genomic characterization of endemic Salmonella enterica serovar Typhimurium and Salmonella enterica serovar I 4,[5],12:i:-isolated in Malaysia. Infect Genet Evol. 2018;62:109–21. https://doi.org/10.1016/j.meegid.2018.04.027.

    Article  CAS  PubMed  Google Scholar 

  26. Figueroa-Bossi N, Bossi L. Inducible prophages contribute to Salmonella virulence in mice. Mol Microbiol. 1999;33(1):167–76. https://doi.org/10.1046/j.1365-2958.1999.01461.x.

    Article  CAS  PubMed  Google Scholar 

  27. Brüssow H, Canchaya C, Hardt WD. Phages and the evolution of bacterial pathogens: from genomic rearrangements to lysogenic conversion. Microbiol Mol Biol Rev. 2004;68(3):560–602. doi:https://doi.org/10.1128/MMBR.68.3.560-602.2004. (Table of contents).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. García P, Malorny B, Hauser E, Mendoza MC, Rodicio MR. Genetic types, gene repertoire, and evolution of isolates of the Salmonella enterica serovar 4,5,12:i:-Spanish clone assigned to different phage types. J Clin Microbiol. 2013;51(3):973–8. https://doi.org/10.1128/JCM.02777-12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Pang S, Octavia S, Feng L, Liu B, Reeves PR, Lan R, et al. Genomic diversity and adaptation of Salmonella enterica serovar Typhimurium from analysis of six genomes of different phage types. BMC Genom. 2013;14:718. doi:https://doi.org/10.1186/1471-2164-14-718.

    Article  CAS  Google Scholar 

  30. Yang X, Wu Q, Zhang J, Huang J, Guo W, Cai S. Prevalence and characterization of monophasic Salmonella Serovar 1,4,[5],12:i:-of food origin in China. PLoS ONE. 2015;10(9):e0137967. https://doi.org/10.1371/journal.pone.0137967.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Asokan GV, Ramadhan T, Ahmed E, Sanad H. WHO global priority pathogens list: a bibliometric analysis of medline-PubMed for knowledge mobilization to infection prevention and control practices in Bahrain. Oman Med J. 2019;34(3):184–93. https://doi.org/10.5001/omj.2019.37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Sun J, Deng Z, Yan A. Bacterial multidrug efflux pumps: mechanisms, physiology and pharmacological exploitations. Biochem Biophys Res Commun. 2014;453(2):254–67. https://doi.org/10.1016/j.bbrc.2014.05.090.

    Article  CAS  PubMed  Google Scholar 

  33. Miryala SK, Ramaiah S. Exploring the multi-drug resistance in Escherichia coli O157:H7 by gene interaction network: a systems biology approach. Genomics. 2019;111(4):958–65. https://doi.org/10.1016/j.ygeno.2018.06.002.

    Article  CAS  PubMed  Google Scholar 

  34. Universal Protein Resource. (UniProt)—resistance genes. https://www.uniprot.org/uniref/?query=resistance+genes&sort=score. Accessed 24 Nov 2020.

  35. Horiyama T, Yamaguchi A, Nishino K. TolC dependency of multidrug efflux systems in Salmonella enterica serovar Typhimurium. J Antimicrob Chemother. 2010;65(7):1372–6. https://doi.org/10.1093/jac/dkq160.

    Article  CAS  PubMed  Google Scholar 

  36. Campioni F, Falcão JP. Genotypic diversity and virulence markers of Yersinia enterocolitica biotype 1A strains isolated from clinical and non-clinical origins. APMIS. 2014;122(3):215–22. https://doi.org/10.1111/apm.12126.

    Article  CAS  PubMed  Google Scholar 

  37. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19(5):455–77. https://doi.org/10.1089/cmb.2012.0021.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gurevich A, Saveliev V, Vyahhi N, Tesler G. QUAST: quality assessment tool for genome assemblies. Bioinformatics. 2013;29(8):1072–5. https://doi.org/10.1093/bioinformatics/btt086.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cao G, Meng J, Strain E, Stones R, Pettengill J, Zhao S, et al. Phylogenetics and differentiation of Salmonella newport lineages by whole genome sequencing. PLoS ONE. 2013;8(2):e55687. https://doi.org/10.1371/journal.pone.0055687.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Papanikolaou N, Trachana K, Theodosiou T, Promponas VJ, Iliopoulos I. Gene socialization: gene order, GC content and gene silencing in Salmonella. BMC Genom. 2009;10:597. https://doi.org/10.1186/1471-2164-10-597.

    Article  CAS  Google Scholar 

  41. Almeida F, Medeiros MIC, Rodrigues DP, Payne J, Timme RE, Allard MW, et al. Draft genome sequences of 40 Salmonella enterica Serovar Typhimurium strains isolated from humans and food in Brazil. Genome Announc. 2018;4(5):e00892-16. https://doi.org/10.1128/genomeA.00892-16.

    Article  Google Scholar 

  42. Seribelli AA, Frazão MR, Gonzales JC, Cao G, Leon MS, Kich JD, et al. Draft genome sequences of 20 Salmonella enterica subsp. enterica Serovar Typhimurium strains isolated from swine in Santa Catarina, Brazil. Genome Announc. 2018;6(16):e00232-18. https://doi.org/10.1128/genomeA.00232-18.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Yu G, Smith D, Zhu H, Guan Y, Lam TTY. ggTree: an r package for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods Ecol Evol. 2017;8:28–36. doi:https://doi.org/10.1111/2041-210X.12628.

    Article  Google Scholar 

  44. Yu G, Lam TT, Zhu H, Guan Y. Two methods for mapping and visualizing associated data on phylogeny using Ggtree. Mol Biol Evol. 2018;35(12):3041–3. https://doi.org/10.1093/molbev/msy194.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Kaas RS, Leekitcharoenphon P, Aarestrup FM, Lund O. Solving the problem of comparing whole bacterial genomes across different sequencing platforms. PloS ONE. 2014;9(8):e104984. https://doi.org/10.1371/journal.pone.0104984.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Larsen MV, Cosentino S, Rasmussen S, Friis C, Hasman H, Marvig RL, et al. Multilocus sequence typing of total-genome-sequenced bacteria. J Clin Microbiol. 2012;50(4):1355–61. https://doi.org/10.1128/JCM.06094-11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Arndt D, Grant JR, Marcu A, Sajed T, Pon A, Liang Y, et al. PHASTER: a better, faster version of the PHAST phage search tool. Nucleic Acids Res. 2016;44(W1):W16–21. https://doi.org/10.1093/nar/gkw387.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. McArthur AG, Waglechner N, Nizam F, Yan A, Azad MA, Baylay AJ, et al. The comprehensive antibiotic resistance database. Antimicrob Agents Chemother. 2013;57(7):3348–57. https://doi.org/10.1128/AAC.00419-13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

We thank São Paulo Research Foundation (FAPESP) (Proc. 2016/24716-3 and Proc. 2019/19338-8) and Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES)—Brasil Finance Code 001 for financial support. During this work, Seribelli, A.A. was supported by a scholarship from São Paulo Research Foundation (FAPESP) (Proc. 2017/06633-6). Falcão, J. P. received a productive fellowship from Council for Scientific and Technological Development (CNPq) Grant 304399/2018-3.

Author information

Authors and Affiliations

Authors

Contributions

AAS participated in the design of the work, performed the methods, analysed and interpreted all the data and wrote the original manuscript. PS performed the methods. MFC analysed and interpreted the data. FA performed the whole genome sequencing of the S. Typhimurium strains isolated from humans and foods. MRF performed the whole genome sequencing of the S. Typhimurium strains isolated from swine. MICM collected S. Typhimurium strains isolated from humans and foods. DPR collected S. Typhimurium strains isolated from foods. JDK collected S. Typhimurium strains isolated from swine. LJB performed the methods. SCS performed the methods. MWA was the WGS supervisor, revised and edited the manuscript. JPF participated in the design of the work, supervised the project, helped to analyze the results obtained and revised and edited the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Marc W. Allard or Juliana Pfrimer. Falcão.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Seribelli, A.A., da Silva, P., da Cruz, M.F. et al. Insights about the epidemiology of Salmonella Typhimurium isolates from different sources in Brazil using comparative genomics. Gut Pathog 13, 27 (2021). https://doi.org/10.1186/s13099-021-00423-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13099-021-00423-7

Keywords