Skip to main content
Advertisement
Browse Subject Areas
?

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

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Genetic diversity of Salmonella enterica isolated over 13 years from raw California almonds and from an almond orchard

  • Anne-laure Moyne,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliations Department of Food Science and Technology, University of California, Davis, California, United States of America, Western Center for Food Safety, University of California, Davis, California, United States of America

  • Opeyemi U. Lawal,

    Roles Formal analysis, Methodology, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Canadian Research Institute for Food Safety, Department of Food Science, University of Guelph, Guelph, Ontario, Canada

  • Jeff Gauthier,

    Roles Formal analysis, Methodology

    Affiliation Institut de biologie intégrative et des systèmes (IBIS), Faculté de médecine, Université Laval, Québec, Québec, Canada

  • Irena Kukavica-Ibrulj,

    Roles Data curation, Formal analysis, Methodology

    Affiliation Institut de biologie intégrative et des systèmes (IBIS), Faculté de médecine, Université Laval, Québec, Québec, Canada

  • Marianne Potvin,

    Roles Formal analysis, Methodology

    Affiliation Institut de biologie intégrative et des systèmes (IBIS), Faculté de médecine, Université Laval, Québec, Québec, Canada

  • Lawrence Goodridge,

    Roles Funding acquisition, Resources, Writing – review & editing

    Affiliations Canadian Research Institute for Food Safety, Department of Food Science, University of Guelph, Guelph, Ontario, Canada, Food Safety and Quality Program, Department of Food Science and Agricultural Chemistry, McGill University, Sainte Anne de Bellevue, Quebec, Canada

  • Roger C. Levesque,

    Roles Funding acquisition, Methodology, Writing – review & editing

    Affiliation Institut de biologie intégrative et des systèmes (IBIS), Faculté de médecine, Université Laval, Québec, Québec, Canada

  • Linda J. Harris

    Roles Conceptualization, Funding acquisition, Project administration, Writing – original draft, Writing – review & editing

    ljharris@ucdavis.edu

    Affiliations Department of Food Science and Technology, University of California, Davis, California, United States of America, Western Center for Food Safety, University of California, Davis, California, United States of America

Abstract

A comparative genomic analysis was conducted for 171 Salmonella isolates recovered from raw inshell almonds and raw almond kernels between 2001 and 2013 and for 30 Salmonella Enteritidis phage type (PT) 30 isolates recovered between 2001 and 2006 from a 2001 salmonellosis outbreak-associated almond orchard. Whole genome sequencing was used to measure the genetic distance among isolates by single nucleotide polymorphism (SNP) analyses and to predict the presence of plasmid DNA and of antimicrobial resistance (AMR) and virulence genes. Isolates were classified by serovars with Parsnp, a fast core-genome multi aligner, before being analyzed with the CFSAN SNP Pipeline (U.S. Food and Drug Administration Center for Food Safety and Applied Nutrition). Genetically similar (≤18 SNPs) Salmonella isolates were identified among several serovars isolated years apart. Almond isolates of Salmonella Montevideo (2001 to 2013) and Salmonella Newport (2003 to 2010) differed by ≤9 SNPs. Salmonella Enteritidis PT 30 isolated between 2001 and 2013 from survey, orchard, outbreak, and clinical samples differed by ≤18 SNPs. One to seven plasmids were found in 106 (62%) of the Salmonella isolates. Of the 27 plasmid families that were identified, IncFII and IncFIB plasmids were the most predominant. AMR genes were identified in 16 (9%) of the survey isolates and were plasmid encoded in 11 of 16 cases; 12 isolates (7%) had putative resistance to at least one antibiotic in three or more drug classes. A total of 303 virulence genes were detected among the assembled genomes; a plasmid that harbored a combination of pef, rck, and spv virulence genes was identified in 23% of the isolates. These data provide evidence of long-term survival (years) of Salmonella in agricultural environments.

Introduction

From 2001 to 2006, Salmonella enterica serovar Enteritidis was implicated in three outbreaks linked to raw almond consumption. Epidemiologic and traceback investigations of a 2000 to 2001 (denoted 2001) salmonellosis outbreak in Canada and in the United States identified a rare phage type (PT) of Salmonella Enteritidis, PT 30, from clinical samples, raw almonds sampled at retail, environmental (swab) samples at an almond processor and a huller-sheller facility, and environmental drag swabs obtained from multiple orchards in California [1]. Salmonella Enteritidis PT 30 was recovered in 2001 (three of six samples collected) [1]) and, in a subsequent study, at least two times in each year between 2002 and 2006 from environmental drag swabs collected in one of these orchards (eight to 96 samples collected each year) [2]. Raw almonds were epidemiologically linked to clinical cases of Salmonella Enteritidis PT 30 reported in Sweden in 2005 to 2006 (denoted 2006) [3] but Salmonella was not isolated from the implicated California almonds. Between 2003 and 2004 (denoted 2004), another rare phage type, Salmonella Enteritidis PT 9c, was linked to consumption of raw California almonds in the United States [4].

Excluding the farms linked to the 2001 outbreak, the prevalence and levels of Salmonella were determined in a multi-year survey of raw almond kernels. Samples of individual lots of raw almond kernels were collected as they were received by handlers (processors) located throughout the almond-growing regions of California from 2001 to 2007, and in 2010 and 2013 [57]. Inshell almonds were included in the survey in 2006 and 2007 [6]. The prevalence of Salmonella in 14,949 lots of raw almond kernels was 0.98% ± 0.29% over the 9 years (146 positive 100-g samples) [7]. Levels of Salmonella were estimated for 118 of the positive lots; mean and median levels of Salmonella were 1.14 ± 1.69 and 0.79 most probable number (MPN)/100 g, respectively, with single lots at 9.25 and 15.4 MPN/100 g [7]. Of the small number of raw inshell almond lots (455) evaluated in 2006 and 2007, 1.5% (seven 100-g subsamples) were positive [6]. Using classical serological nomenclature, the Salmonella isolates retrieved from these surveys were serotyped into 45 different serovars, including Salmonella Enteritidis PT 30 and PT 9c [5, 6].

Pulsed-field gel electrophoresis (PFGE), multilocus variable-number tandem repeat analysis (MLVA), and comparative genomic indexing (CGI) were applied to Salmonella Enteritidis strains associated with the 2001, 2004, and 2006 almond outbreaks, including clinical and environmental isolates [8]. The Salmonella Enteritidis PT 30 and PT 9c strains could be separated from each other and from other Salmonella Enteritidis phage types based on DNA enzyme restriction profiles, MLVA types, and genes identified by CGI. However, neither PFGE nor MVLA could discriminate among the Salmonella Enteritidis PT 30 isolates associated with the 2001 and 2006 almond-associated outbreaks. Salmonella Enteritidis PT 30 almond-associated outbreak strains could not be distinguished from epidemiologically unrelated Salmonella Enteritidis PT 30 clinical strains included in the study [8]. Among Salmonella Enteritidis PT 30 strains with identical genotypes, metabolic analyses with Biolog revealed differences between clinical and environmental isolates [8], indicating the discriminatory limit of the then current genotyping methods. Environmental Salmonella Enteritidis PT 30 isolates, collected between 2001 and 2006 from one of the 2001 outbreak-associated orchards, clustered in two groups based on the separation by PFGE of their XbaI-digested DNA [2].

Whole genome sequencing (WGS) has replaced PFGE for investigating foodborne outbreaks because of its higher resolution, and this methodology has been incorporated into routine public health surveillance since 2014 [9, 10]. Clonality of pathogen strains determined by WGS analyses can provide information about contamination during food production and distribution [11]. Different workflows have been developed to assess levels of genetic relatedness of foodborne pathogens by the National Center for Biotechnology Information (NCBI) with the Pathogen Detection Pipeline (https://www.ncbi.nlm.nih.gov/pathogens/), the U.S. Food and Drug Administration with the Center for Food Safety and Applied Nutrition CFSAN SNP Pipeline [12], and the U.S. Centers for Disease Control and Prevention with Lyve-SET [13], based on the specific needs of the respective agencies. The CFSAN SNP Pipeline creates high quality SNP matrices from WGS data that allow connection between clinical isolates to food or environmental isolates based on their evolutionary relationship [12]. Resident pathogen strains in facilities will have closely related WGS profiles, whereas transient pathogen strains will have unique or unrelated WGS profiles [11, 14].

The objective of the present study was a comparative genomic analysis of 171 Salmonella (45 serovars) isolated from raw inshell almonds and almond kernels in 9 years of surveys conducted between 2001 and 2013; 30 Salmonella Enteritidis PT 30 isolates recovered between 2001 and 2006 from a 2001 outbreak-associated almond orchard were included in the analysis. The CFSAN SNP Pipeline was selected to measure the genetic distances among the isolates. WGS was used to predict antimicrobial resistance (AMR), virulence genes, and presence of plasmid DNA.

Materials and methods

Isolate selection

Isolates were retrieved from enrichment of raw almond kernels and inshell almonds (survey isolates), and from swabs that were dragged across the floor of one of the 2001 outbreak-associated orchards (orchard isolates). A total of 15,505 ∼1-kg samples from different lots of raw almond kernels (14,949) and inshell almonds (455) were collected upon receipt at several almond processors located throughout California from the 2001–2005 [5], 2006–2007 [6], 2010 ([15]; present study), and 2013 ([7]; present study) harvests, as described previously. All samples were coded to keep the identities of the processors confidential; the geographic origins of the samples were unknown. The Safe Food Alliance, Kingsburg, CA (formerly American Council for Food Safety and Quality, Fresno, CA) analyzed a 100-g subsample from each lot of almonds by enriching for the presence of Salmonella [5]. In addition, the MPN of Salmonella was determined for 118 samples by enriching one or more additional samples weighing from ∼0.25 to 100 g. Salmonella isolates were stored at −80°C.

A total of 171 Salmonella survey isolates were selected for the present study: for each individual positive almond lot, a single Salmonella isolate retrieved from the initial 100-g subsample (153 positive samples, 148 isolates; three, one, and one isolate from 2001, 2002, and 2003, respectively, were lost), any isolates recovered from MPN or secondary enrichments that differed from the initial serotype (12), and unique isolates recovered from secondary enrichments of initially negative samples (11) (Tables 1 and S1 in S1 File). Traditional serotyping was done for each banked Salmonella by the California Animal Health and Food Safety Laboratory System (Davis, CA), and phage typing for Salmonella Enteritidis isolates was done by the National Veterinary Services Laboratory (Ames, IA) (Table 1). Salmonella Enteritidis PT 9c LJH1024 (obtained from Robert Mandrell, U.S. Department of Agriculture, Agricultural Research Service; RM4635 or G04-101 [8], raw almond isolate from the 2004 outbreak) and Salmonella Enteritidis PT 30 LJH0608 (raw almond isolate from the 2001 outbreak that was deposited to the American Type Culture Collection as ATCC BAA-1045) were also included in some of the analyses (S2 Table in S1 File).

thumbnail
Table 1. Results of the WGS serotyping and number of plasmids for Salmonella survey isolates retrieved from raw almonds.

https://doi.org/10.1371/journal.pone.0291109.t001

Several 2001 outbreak-associated almond orchards sampled by investigators during the 2001 outbreak investigation were positive for Salmonella Enteritidis PT 30 [1]. One of these orchards was sampled every year from 2001 [1] through 2006 [2]. Briefly, sterile gauze swabs attached to a string and soaked in full-strength evaporated skim milk were pulled along the orchard floor in a standardized manner. Four individual swabs were pooled and a procedure designed for recovering Salmonella from poultry houses was used to enrich the samples. Three of six (50%) pooled swab samples collected in 2001 and 53 of 228 (23%) samples collected between 2002 and 2006 were positive for Salmonella; every isolate was identified as Salmonella Enteritidis PT 30. A total of 30 Salmonella Enteritidis PT 30 orchard isolates from 2001 (3), 2002 (12), 2003 (10), 2005 (2), and 2006 (3) were analyzed in the present study (S2 Table in S1 File). One orchard isolate from 2002 and all orchard isolates from 2004 (25) were not available and thus not included.

Using the pathogen detection tool associated with the NCBI database (https://www.ncbi.nlm.nih.gov/pathogens), the sequence read archive (SRA) data were downloaded for 12 Salmonella Enteritidis PT 30 clinical isolates from the 2001 almond outbreak and for four clinical isolates from the 2006 almond outbreak (S3 Table in S1 File).

Whole genome sequencing

Isolates were retrieved from frozen glycerol stock and plated on tryptic soy agar (TSA). Following overnight incubation at 37°C, one colony was inoculated into 2 ml of tryptic soy broth (TSB) and incubated for 24 h at 37°C, with shaking at 120 rpm, before being pelleted by centrifugation at 14,000 × g for 2 min. Genomic DNA, for the isolates in S2 Table (S1 File), was extracted with the QIAamp DNA minikit (Qiagen, Valencia, CA) following the manufacturer’s directions. The 150-bp paired-end libraries were constructed for each purified DNA with the Illumina Nextera DNA flex library kit following the manufacturer’s directions (Illumina Inc., San Diego, CA). Pooled samples were sequenced on an Illumina 4000 HiSeq system by the DNA Technologies and Expression Analysis Core at the UC Davis Genome Center. DNA, for all the survey isolates, was sequenced as described by Emond-Rheault et al. in 2020 [16]. All sequence data obtained in this study were deposited to the NCBI pathogen database under the BioProject accession number PRJNA941918 (Tables 1 and S2 in S1 File) and PRJNA951760 (S2 Table in S1 File).

Quality control and genome assembly

Raw read quality was assessed with FastQC (v0.11.8) (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/). Illumina adapter sequences and low-quality sequences were trimmed using Trimmomatic version 0.36 [17]. Reads were assembled de novo with SPAdes version v3.13.0 Genome Assembler [18]. Draft Salmonella genome assemblies were serotyped with SeqSero [19] and by aligning with BLASTn against the nonredundant nucleotide sequence database.

Core genome SNP typing

De novo genomes were used to build core genome single nucleotide trees using the Parsnp aligner v1.2 with default parameters and the requirement for all genomes to be included in the analysis [20]. The 171 Salmonella genomes from the survey were mapped to the complete reference genome Salmonella Typhimurium str. LT2 (NCBI accession: AE006468), resulting in an alignment of 50% of the core genome. The whole-genome phylogeny was constructed with FastTree2 [21]. iTOL (http://itol.embl/de/) was used to visualize the tree and annotate it with the AMR genes [22].

Genetic distance

Genetic distance between multiple isolates of the same serovar was evaluated as the number of SNP differences detected with the CFSAN SNP Pipeline [12]. The CFSAN SNP Pipeline v2.0.2 was installed on a local ubuntu platform with all the executable software dependencies. Prior to analyzing our data, we used the data set provided for testing the reproducibility of the software to confirm that our installation of the CFSAN SNP Pipeline was correct. Reference-based alignments were created for a set of samples and used to generate the SNP matrix. Because the software was developed for closely related genome sequences, where available, complete assembled reference genomes were downloaded from NCBI (20; S4 Table in S1 File). The phylogenies were inferred with MEGA7 [23] using the neighbor-joining method [24] based on the obtained SNP matrix formatted as a FASTA file generated by the CFSAN SNP pipeline v2.0.2.

Virulence and antibiotic resistance (AMR) gene prediction

The presence of resistance genes, as well as point mutations, were determined using ResFinder 4.1 (Center for Genomic Epidemiology, https://cge.food.dtu.dk/services/ResFinder) with a setting threshold of 90% and minimum length of 60% [2527]. The assembled draft genomes for the survey isolates (171) were used as an input to identify Salmonella AMR genes associated with resistance to aminoglycoside, β-lactam, chloramphenicol, colistin, fluoroquinolone, fosfomycin, glycopeptide, macrolide, sulfonamide, tetracycline, and trimethoprim antibiotics. The VF analyzer pipeline was used to screen the assembled draft genomes against the Virulence Factor Database (VFDB) for potential virulence factors [28].

Plasmid detection and reconstruction

Plasmids from genome assemblies were typed and reconstructed using MOB-suite v3.1.0 with the default parameters [29, 30]. To determine the AMR and virulence genes that were plasmid-borne, the reconstructed plasmids were screened against the CARD (https://card.mcmaster.ca) and VFDB [28] databases, respectively, using Abricate v0.5 (https://github.com/tseemann/abricate) with the same parameters as described above.

Results

Phylogenetic analysis of Salmonella isolates retrieved from raw almonds

Genomes were initially compared with Parsnp because the CFSAN SNP Pipeline is not recommended for relatively distant bacteria (greater than a few hundred SNP differences). A maximum-likelihood phylogeny tree was constructed with Parsnp based on alignment of the 171 Salmonella assembled genomes [20]. All the isolates belonged to the species enterica, with most (165) belonging to the subspecies enterica (Fig 1 and Table 1). Three isolates were classified as subspecies diarizonae and three as subspecies arizonae. The phylogenetic tree clustered the isolates by serovar (Fig 1). The serovars identified by classical serotyping matched the serovars predicted by WGS for 92% of the isolates (Table 1). However, 14 isolates clustered with serotypes that differed from those to which they were initially assigned by traditional serotyping (Table 1). To eliminate the possibility of manipulation errors, these isolates were resequenced and their serotype was confirmed with SeqSero and with BLASTn against the nonredundant nucleotide sequence database. Results were consistent with the initial WGS serovar prediction.

thumbnail
Fig 1. Maximum likelihood tree based on SNPs identified by aligning 171 de novo assemblies to the reference chromosome of Salmonella Typhimurium LT2 ASM694v2 with Parsnp and genotypic antimicrobial resistance (AMR).

The scale bar is in number of SNPs. The colors in the phylogenetic tree represent different serogroups; squares represent a chromosomal location and circles represent a plasmid location for AMR genes. AMR genes are color-coded by antibiotic classes.

https://doi.org/10.1371/journal.pone.0291109.g001

Based on the core genome SNP typing, a total of 32 unique Salmonella serovars were identified. Of these, 22 Salmonella serovars were isolated two or more times between 2001 and 2013: Enteritidis (n = 16), Muenchen (n = 16), Newport (n = 15), Montevideo (n = 14), Typhimurium (n = 12), Thompson (n = 11), Give (n = 10), 1,4,[5],12:i- (n = 9), Oranienburg (n = 7), Heidelberg (n = 6), Infantis (n = 5), Senftenberg (n = 5), Anatum (n = 4), Cerro (n = 4), Brandenburg (n = 3), Duisburg (n = 3), Horsham (n = 3), Manhattan (n = 3), Tennessee (n = 3), Agona (n = 2), Braenderup (n = 2), and Mbandaka (n = 2) (Fig 1 and Table 1). Three isolates initially identified as Salmonella Othmarschen (LJH0752, LJH1143, and LJH1276-1) clustered with Salmonella Oranienburg. Both serotypes have a similar antigenic formula, 6,7,14:m,t:-, which makes them difficult to distinguish by serological methods [31]. Serovar 1,4,[5],12:i:- is a monophasic variant of Salmonella Typhimurium, and three of 14 isolates initially identified as Typhimurium (LJH0738, LJH1662-1, LJH1676) clustered with 1,4,[5],12:i:-. One Salmonella initially identified as 1,4,[5],12:i:- (LJH1144) clustered with Typhimurium.

Among the five isolates that were untypeable by classical serotyping, two clustered with serovar Muenchen (LJH1151, LJH1620-1), one with Enteritidis (LJH1029), and one with Manhattan (LJH1026). A single untypeable isolate (LJH1624-1) clustered with Salmonella diarizonae (LJH1664) and was identified as Salmonella diarizonae with BLASTn and SeqSero [19]. Fifteen unique serotypes, predicted with BLASTn and SeqSero, matched the serological serotyping. It was difficult to predict some serovars due to the limited number of representative genomes in the SeqSero database. Salmonella Zerifin LJH0724 was identified as Salmonella Istanbul with SeqSero and clustered with the Istanbul isolates in the Parsnp phylogeny tree (Fig 1). Conflicting serotyping results between traditional methods and WGS have been reported in previous studies [3235]. Serotyping based on WGS has been increasingly used by public health laboratories and federal agencies to replace the current standard of phenotypic serotyping [10, 32, 34, 36]. In the present study, serotype was assigned based on the WGS analysis when the serovar assignment was identical among the phylogenetic relationship, SeqSero, and BLASTn analyses. Except for a single strain of Salmonella Zerifin, all isolates (170 of 171) were assigned a serotype based on the WGS analysis.

Genetic distance within each serovar

The CFSAN SNP Pipeline was used to evaluate the genetic distance among multiple isolates within each serotype. Salmonella Enteritidis isolates clustered in three groups by phage type: PT 8, PT 9c, and PT 30 (Fig 2). All five Salmonella Enteritidis PT 8 isolates were retrieved in 2005 from separate almond lots and had less than three SNP differences, classifying them as clonal isolates. Salmonella Enteritidis PT 9c isolated in 2005 (LJH1028) and 2010 (LJH1272), differed by 5 and 13 SNPs, respectively, from a 2004 Salmonella Enteritidis PT 9c outbreak isolate (LJH1024; Fig 2).

thumbnail
Fig 2. Phylogenetic tree of Salmonella Enteritidis generated with the CFSAN SNP Pipeline.

The branch lengths represent the SNP distances among the isolates. Almond survey isolates (black text) and outbreak-associated orchard isolates (green dot) were compared to clinical isolates (red text) from the 2001 and 2006 outbreaks (retrieved from the NCBI database). Almond isolates of Salmonella Enteritidis PT 9c (LJH1024) and Salmonella Enteritidis PT 30 (LJH0608) from the 2004 and 2001 outbreaks, respectively, were included for comparison.

https://doi.org/10.1371/journal.pone.0291109.g002

The genomes of Salmonella Enteritidis PT 30 recovered from survey almonds (eight isolates (Table 1): LJH0762 [2003], LJH1023 [2005], LJH1104 [2006], LJH1059 [2006], LJH1096 [2006], LJH1109 [2006], LJH1633 [2013], LJH1673 [2013]), the 2001 outbreak-associated orchard (30 isolates; S2 Table in S1 File), and a 2001 outbreak-associated almond isolate (LJH0608) were compared to Salmonella Enteritidis PT 30 genomes of clinical isolates from almond outbreaks in 2001 (12 isolates) and 2006 (four isolates) (S3 Table in S1 File).

Salmonella Enteritidis PT 30 isolates formed two clusters (Fig 2). One consisted of a single survey isolate (LJH0762), recovered in 2003, that differed from LJH0608 by 48 SNPs (Fig 2 and S5 Table in S1 File). All other survey and clinical isolates (n = 38) clustered in a single group with LJH0608 that differed from each other by ≤18 SNPs (Fig 2) indicating that the isolates are from a common origin. Almond isolates from 2001 to 2013 had 2 to 13 SNP differences compared with the 2001 outbreak-associated almond isolate Salmonella Enteritidis PT 30 LJH0608 (Fig 2 and S5 Table in S1 File). Although this isolate was recovered from recalled almonds in 2001, the almonds were harvested in the fall of 2000 [1], a span of 14 years (2000–2013). The orchard isolates from 2001 to 2006 differed by 0 to 12 SNPs within their genomes and by 3 to 13 SNPs with the clinical genomes. The SNP differences ranged from zero to eight within the 12 clinical isolates from the 2001 outbreak and from one to 13 within the four clinical isolates from the 2006 outbreak. Among the clinical isolates from 2001 and 2006, the SNP differences ranged from four to 13, indicating that the isolates are from a common origin.

Almond, orchard, and clinical isolates of Salmonella Enteritidis PT 30 isolated from 2001 through 2013 are closely related strains. The persistence of Salmonella Enteritidis PT 30 in an almond orchard over 6 years was reported previously [2]. The SNP analysis confirmed the PFGE results obtained for these isolates. Almonds from the 2001 outbreak-associated orchards were purposefully excluded from the raw almond survey. Because survey samples were coded and the sources unknown, it is possible that samples harvested from outbreak-associated orchards were inadvertently included. It is also possible that survey almonds were cross contaminated with almonds harvested from outbreak-associated orchards via harvest equipment or at a common almond huller-sheller, or that Salmonella Enteritidis PT 30 was spread over a broader geographic region than recognized as associated with the 2001 outbreak.

Salmonella Enteritidis was not the only serovar for which clonal isolates were recovered from almonds in different years. Salmonella Montevideo survey isolates clustered into three groups, with more than 100 SNPs between them (Fig 3). Within each of these clusters there were isolates separated by one or more years (including isolates from 2001 and 2013) that differed by ≤9 SNPs, indicating that they share a common ancestor. Salmonella Newport (S6 Table in S1 File) and Salmonella Muenchen (S7 Table in S1 File) each clustered into two groups separated by more than 100 SNPs. A small number of SNP differences (<3) between Salmonella Newport genomes (S6 Table in S1 File) were identified in isolates retrieved in 2003 (LJH0751), 2006 (LJH1084), 2007 (LJH1133), 2010 (LJH1248, LJH1273 and LJH1278). Nine isolates of Salmonella Muenchen were retrieved in 2006; eight of these isolates from six different almond lots had nearly identical genomes, with ≤3 SNP differences (S7 Table in S1 File), and differed from single isolates in 2007 (LJH1135) and 2013 (LJH1661) by ≤8 SNPs. Closely related genomes for single isolates of Salmonella Anatum (S8 Table in S1 File) and Salmonella Thompson (S9 Table in S1 File) were identified 2 and 1 years apart, respectively. Several Salmonella Oranienburg (S10 Table in S1 File) and Salmonella Typhimurium (Fig 4) were identified 5 and 3 years apart, respectively. Closely related genomes (≤13 SNPs) were identified for isolates retrieved from separate almond lots during the same year for Salmonella serovar Cerro (S11 Table in S1 File), Give (S12 Table in S1 File), Infantis (S13 Table in S1 File), Heidelberg (S14 Table in S1 File) and Tennessee (S15 Table in S1 File). For Salmonella serovar Brandenburg (S16 Table in S1 File), Braenderup, Manhattan (S17 Table in S1 File), Mbandaka, and Senftenberg (S18 Table in S1 File), isolates were separated by more than 13 SNPs.

thumbnail
Fig 3. Phylogenetic tree of Salmonella Montevideo generated with the CFSAN SNP Pipeline.

The branch lengths represent the SNP distances among the isolates.

https://doi.org/10.1371/journal.pone.0291109.g003

thumbnail
Fig 4. Phylogenetic tree of Salmonella Typhimurium and 1,4,[5],12:i:-.

The branch lengths are representative of the SNP distances among the isolates.

https://doi.org/10.1371/journal.pone.0291109.g004

Multiple SNP-based approaches have been developed to analyze the large number of short reads produced by various sequencing platforms [37, 38]. The CFSAN Pipeline was selected because it is less sensitive to coverage changes [37] and has good discriminative power. Its high resolution, however, strongly depends on an appropriate reference genome. In addition to outbreak source attribution, this tool can reveal similarity among isolates and their persistence in food facilities or in the production environment [14, 39].

Differences of more than 100 SNPs were sometimes observed among isolates within each serovar. Survey isolates that were genetically similar (≤13 SNPs) were recovered in multiple years, up to 13 years for Salmonella Enteritidis PT 30 and Salmonella Montevideo and 10 years for Salmonella Newport. Because the survey samples were collected after hulling and shelling and prior to entering a processing facility or storage, the contamination source would have to be at production or harvest (orchard), during postharvest handling (transportation to huller, during storage, or during hulling and shelling), or transportation from huller to processor. Although the exact geographic locations of the almond survey samples were unknown, clonal strains of Salmonella Enteritidis PT 8 (five; 2005) and Salmonella Muenchen (eight; 2006) were isolated from different lots of almonds in the same year.

The diversity of serovars and unique strains within serovars was expected in a survey that likely reflects broad environmental contamination (e.g., almond orchard during production or harvest). Common harvest and postharvest practices may also lead to distribution of Salmonella in almonds from geographically diverse orchards that share the same equipment or facilities. At maturity, almonds are shaken to the ground where they dry for several days. They are then harvested by windrowing and sweeping off the orchard floor. The in-hull, inshell almonds are then transported to facilities where the hull and shell are removed. Kernels mix with hulls and shells before sorting, separation, and bulk transportation to processing facilities. Once kernels are delivered to almond processing facilities, commingling of almond lots may occur prior to or during storage. These practices may explain the clusters of Salmonella Enteritidis PT 8 and Salmonella Muenchen isolated from different lots in 2005 and 2006.

U.S. regulations were implemented in 2007 that require all California-grown almonds sold in North America (U.S., Canada, and Mexico) to be processed with a treatment capable of achieving a minimum 4-log reduction in Salmonella [40]. While there have been outbreaks associated with almond-containing products such as blended nut butters, none have been associated with contaminated almonds since 2006, likely due to effective implementation of these regulations [7, 15].

The persistence of Salmonella has been described for other pre- and postharvest scenarios [11, 41, 42]. A narrow range of Salmonella serovars has been associated with California pistachio outbreaks, outbreak investigations, and industry and retail surveys [41, 43]. Pistachio-associated isolates of Salmonella Senftenberg and Salmonella Montevideo recovered over multiple years (2009 to 2017) and from multiple facilities differed (within each serovar) by 0 to 31 SNPs, and the authors suggested that the organisms may have established residence in the primary production environment or orchards [41].

Plasmid prediction and characterization

Plasmids contribute significantly to the emergence and spread of genes encoding AMR, virulence, and other metabolic functions in multiple scales across Salmonella serotypes [16, 44]. Plasmid carriage in the Salmonella strains under study were assessed by screening and reconstructing the plasmid sequences from assembled genomes using the clustered plasmid reference database-based pipeline [29, 30]. A total of 106 of the 171 Salmonella isolates (62%) carried one to seven plasmids (total plasmids 161; Table 1) with sizes of 1,030 bp to 303,322 bp (Fig 5A and S19 Table in S1 File). Using the presence of relaxase and mate-pair formation marker genes and/or oriT sequence, most of these plasmids were predicted to be either mobilizable (n = 51/161; 32%) or conjugative (n = 83/161; 52%) (Fig 5A). In total, 61 plasmid clusters with 27 different plasmid families were identified, with IncFII and IncFIB being the most predominant types in the collection (Fig 5B).

thumbnail
Fig 5. General features of plasmids identified in Salmonella survey isolates included in this study.

(A) Putative plasmids predicted were categorized as mobilizable, conjugative, or non-mobilizable based on the presence or absence of relaxase, mate-pair formation marker genes and/or oriT sequence. Each colored circle represents a plasmid-carrying Salmonella isolate. (B) Frequency of plasmid replication types.

https://doi.org/10.1371/journal.pone.0291109.g005

Thirty eight Salmonella strains distributed across six serovars (Duisburg, Enteritidis, Lomalinda, Muenchen, Typhimurium, and 1,4,[5],12:i:-) carried plasmids that contained two to seven virulence genes, but no AMR genes (S20 Table in S1 File). The IncFIB-IncFII plasmid family combination was common among these strains (Fig 6), predominantly in Salmonella Enteritidis (n = 15), Typhimurium (n = 8), and 1,4,[5],12:i:- (n = 9). Comparative analysis of the plasmid sequences with known plasmids using mash [45] and BLASTn revealed that these plasmids are similar (nucleotide sequence homology = 100%) to plasmid pCFSAN076214_2 (accession number CP033342.1) described previously in Salmonella Enteritidis strain ATCC BAA-1045 isolated from raw almonds (also LJH608 in this study) [46] and to plasmid p11-0972.1 (accession number CP039855.1) reported in S. enterica serovar 1,4,[5],12:i:- recovered from a human stool sample [47]. IncF plasmids are among the most common plasmids found in Salmonella and are reported to carry multiple antibiotic resistance and/or virulence genes, suggesting their role in the dissemination of these genes across Salmonella serotypes and Enterobacteriaceae by extension [48, 49].

thumbnail
Fig 6. Heatmap of the distribution of plasmid replication types detected in Salmonella isolates.

https://doi.org/10.1371/journal.pone.0291109.g006

Thirteen Salmonella strains belonging to seven serotypes (Agona, Anatum, Heidelberg, Istanbul, Newport, Typhimurium, and Zerifin) carried plasmids with at least one AMR gene but no virulence genes. The most common among these was within the IncC plasmid family and was associated with six to 10 AMR genes (S20 Table in S1 File). The IncC plasmid identified in the present study was predicted to be conjugative, predominant in Salmonella Anatum and Salmonella Newport (Fig 6) and was almost indistinguishable (99.99% by BLASTn) from Salmonella Anatum plasmid pSAN1-1736 (accession number: CP014658.1) [50]. IncC plasmids are known to be widely distributed across Salmonella serotypes from diverse sources and often carry multiple AMR genes [49].

A unique Salmonella Kentucky strain LJH1044 carried a large conjugative plasmid (146 Kb) with IncFIB-IncFIC-rep_cluster_2244 plasmid replicons and contained multiple virulence and AMR genes (Fig 6 and S20 Table in S1 File). This plasmid carried three AMR genes encoding resistance to aminoglycoside and tetracycline, a complete iroBCDEN operon that encodes salmochelin siderophore [51], and an iucABCD-iutA operon that has been described to be associated with aerobactin synthesis essential for virulence and stress response in Salmonella [52] (S20 Table in S1 File). Comparative analysis showed that this plasmid had 99.99% nucleotide sequence similarity to plasmid pCVM29188_146 (accession number CP001122.1) reported previously in several Salmonella Kentucky strains from poultry in the United States [53, 54].

Antimicrobial resistance profile

Among the isolates retrieved from the almond survey, a total of 24 AMR genes were identified with the ResFinder and CARD database, which classified into nine different antimicrobial protein groups: aminoglycosides, β-lactams, colistin, fosfomycin, glycopeptide, phenicol, sulfonamide, tetracycline, and trimethoprim. One gene, aac(6’)-Iaa, which confers resistance to aminoglycosides, was detected in the chromosome of all Salmonella subspecies diarizonae (3) and enterica (165) but not in arizonae (3) (Fig 1). The frequency of Salmonella enterica subspecies enterica isolates that carried an aminoglycoside acetyltransferase aac(6’)-Iaa has been reported to be high (>95%) in multiple WGS analysis [5557]. However, in a surveillance study of non-typhoidal Salmonella enterica, 11 isolates out of 3,491 (0.3%) showed phenotypic resistance to an aminoglycoside antimicrobial [55].

The other 23 AMR genes were detected in 35 isolates, with fosA7 being the most common (Fig 1). The gene fosA7, which confers resistance to fosfomycin, a broad-spectrum cell wall synthesis inhibitor, was first identified in the chromosome of Salmonella Heidelberg isolated from broiler chickens [58]. In the present study, all Salmonella Heidelberg isolates (n = 6), and some isolates of Salmonella serovars Agona (n = 2), Meleagris (n = 1), Montevideo (n = 8), Oranienburg (n = 2), and Tennessee (n = 3), carried the chromosomal fosA7 gene. Ten aminoglycoside resistance genes were detected in 16 isolates: aac(6’)-Iaa, aadA2, aadA3, aadA7, aadA12, aadA13, ant(2”)-Ia, aph(3")-IIa, aph(3’’)-Ib, aph(6)-lc, and aph(6)-Id. The blaCMY-2 and blaCARB-2 genes, which confer resistance to β-lactams, and the floR gene, which confers resistance to phenicol, were found in nine isolates. Three tetracycline efflux resistance genes were identified in 16 isolates: tetA, tetB, and tetG. Sulfisoxazole resistance, encoded by sul1 or sul2, was detected in 11 isolates. The dihydrofolate reductase resistance gene, dfrA12, which confers resistance to trimethoprim, was detected in one Salmonella Newport isolate. The plasmid-mediated colistin resistance and phosphoethanolamine transferase mcr-9.1 gene was detected in one Salmonella Agona isolate. Resistance to bleomycin, encoded by the bleomycin-binding protein gene (ble), was predicted for one Salmonella Heidelberg isolate. All the isolates, except Salmonella serovars Enteritidis, Irumu, Lomalinda, Typhimurium, and 1,4,[5],12:i-, contained a missense mutation in parC associated with resistance to quinolone. Multiple mutations in the quinolone resistance determining region are usually required to confer resistance to ciprofloxacin, but one mutation confers resistance to nalidixic acid [59]. One Salmonella Senftenberg isolate had a point mutation in the 16S rRNA that is thought to confer resistance to spectinomycin.

Most of the predicted AMR genes were identified in a small number (16 out of 171; 9%) of survey isolates and were plasmid encoded in 11 of 16 cases (Fig 1). Multidrug-resistant isolates (putative resistance to at least one antibiotic in three or more drug classes; https://www.cdc.gov/narms/resources/glossary.html) were identified among Salmonella serovars Agona (LJH1100 and LJH1106), Anatum (LJH0720, LJH0787, LJH1012, and LJH1063), Heidelberg (LJH1622), Newport (LJH0656, LJH1138, and LJH1107), and Typhimurium (LJH0788 and LJH1141) (Fig 1). Antibiotic resistance profiles by the calibrated dichotomous sensitivity method were determined for Salmonella isolated from 2001 through 2005 but not for isolates from 2006, 2007, 2010, and 2013. Ten of the Salmonella survey isolates from 2001 to 2005 were resistant to three or more antibiotics [5]. Resistance genotype and phenotype correlated highly for five of these isolates: Salmonella Anatum (n = 3), Salmonella Istanbul (n = 1), and Salmonella Typhimurium var. Copenhagen (n = 1).

Analysis of virulence factors

Salmonella serovars infect a wide range of hosts with different degrees of disease severity, with Enteritidis, Newport, Typhimurium, Javiana, and 1,4,[5],12:i- being significantly more likely to cause illness in humans in the United States [60]. Differences in virulence factors contribute to the severity and outcome of salmonellosis and can be specific to serovars [60]. The Virulence Factor Database (VFDB) was used to detect a total of 303 virulence genes among the 171 Salmonella assembled genomes (Figs 79). Genes were classified under major virulence factors, including the secretion system, fimbrial and non fimbrial adherence, macrophage inducible genes, magnesium uptake, serum resistance, stress proteins, toxins, immune invasion, and two component regulatory systems. The Salmonella pathogenicity island 1 (SPI-1) and 2 (SPI-2), responsible for the type III secretion system, are ubiquitous in S. enterica subsp. enterica [60] and were common to all 171 survey isolates (Fig 7).

thumbnail
Fig 7. Heatmap of the distribution of virulence genes encoding type III secretion system (T3SS) across 171 genomes.

The shades of gray represent the number of open reading frames (ORF) that are detected in each putative virulence gene.

https://doi.org/10.1371/journal.pone.0291109.g007

thumbnail
Fig 8. Heatmap of the distribution of genes encoding the fimbrial operon across 171 genomes.

The shades of gray represent the number of open reading frames (ORF) that are detected in each putative virulence gene.

https://doi.org/10.1371/journal.pone.0291109.g008

thumbnail
Fig 9. Heatmap of the distribution of other virulence genes across 171 genomes.

The shades of gray represent the number of open reading frames (ORF) that are detected by in each putative virulence gene.

https://doi.org/10.1371/journal.pone.0291109.g009

Among the fimbrial adherence factors, the genes encoding the curli fimbriae, csgA, csgB, and csgE, the bcfABCDEFG operon, and the genes that encode type 1 fimbriae, fimA, fimB, fimC, fimD, fimH, fimI, fimW, and fimZ, were also present in all the isolates (Fig 8). The two- component regulatory system phoP-PhoQ genes and the magnesium uptake genes, mgtB-mgtC, (part of SPI-3) were present in all the isolates (Fig 9). All the isolates had the microphage inducible gene, mig-14, but mig-5 was mainly detected in isolates of Salmonella serovars Enteritidis, Typhimurium, and 1,4,[5]12:I:- (Fig 9).

The typhoid toxin genes, cdtB and pltA, originally identified in serotype Typhi, were found in all Salmonella serovar Brandenburg, Duisburg, Horsham, Montevideo, Sandiego, and Schwartzengrund isolates and in nine of 10 Salmonella Oranienburg isolates (Fig 9). However, pltB, required for forming holotoxin, was not present in any of these isolates. In Salmonella Horsham isolates, two genes were identified as homologs of the enterotoxin hemolysin genes of Escherichia coli (hylE/clyA) (Fig 9).

A virulence plasmid that harbored a combination of pef, rck, and spv virulence genes was identified in 23% of the isolates (S20 Table in S1 File). The assembly of the major Pef fimbriae depends on the pefBACDorf5orf6 operon, which encodes PefA fimbriae subunit, PefC usher protein, and the pefD periplasmic chaperone [61]. Eight of 10 Salmonella Typhimurium and all nine Salmonella 1,4,[5]12:I:- isolates carried a plasmid encoding pefA, pefB, pefC, and pefD genes (Fig 8 and S20 Table in S1 File).

The spv genes play a role in suppression of the innate immune response and are often associated with invasive disease and increased virulence [62, 63]. The spv genes were detected in all Salmonella serovar Lomalinda, Enteritidis, and 1,4,[5]12:I:- isolates, in eight of 10 Salmonella Typhimurium isolates, and one of 10 Salmonella Give isolates (Fig 9 and S20 Table in S1 File). The rck gene, which provides protection against the complement-mediated immune response of the host [64], was found in one of 10 Salmonella Give isolates, in all Salmonella Enteritidis PT 30 and PT 8 isolates, in eight of 10 Salmonella Typhimurium isolates, and all Salmonella 1,4,[5]12:I;- isolates (Fig 9).

This study provides one of the first in-depth longitudinal characterizations of Salmonella strains isolated from a single product (almonds) or production environment (almond orchard), in a single geographical region (Central California). This isolate collection is important for understanding Salmonella populations in a significant food production region of the United States. Several clonal strains of Salmonella were isolated over multiple years, adding to a growing body of evidence that enteric pathogens may persist over long periods of time (years) in agricultural environments and in postharvest or food processing facilities (https://www.cdc.gov/ncezid/dfwed/outbreak-response/rep-strains.html) [41, 42]).

Acknowledgments

We thank Sylvia Yada for editing the manuscript and Vanessa Lieberman for technical assistance.

References

  1. 1. Isaacs S, Aramini J, Ciebin B, Farrar JA, Ahmed R, Middleton D, et al. An international outbreak of salmonellosis associated with raw almonds contaminated with a rare phage type of Salmonella Enteritidis. J Food Prot. 2005;68: 191–198. pmid:15690826
  2. 2. Uesugi AR, Danyluk MD, Mandrell RE, Harris LJ. Isolation of Salmonella Enteritidis phage type 30 from a single almond orchard over a 5-year period. J Food Prot. 2007;70: 1784–1789. pmid:17803132
  3. 3. Ledet Muller L, Hjertqvist M, Payne L, Pettersson H, Olsson A, Plym-Forshell L, et al. Cluster of Salmonella Enteritidis in Sweden 2005–2006 –suspected source: almonds. Eurosurveillance. 2007;12: 9–10. pmid:17991404
  4. 4. CDC Centers for Disease Control and Prevention. Outbreak of Salmonella serotype Enteritidis infections associated with raw almonds—United States and Canada, 2003–2004. MMWR Morb Mortal Wkly Rep. 2004: 53: 484–487. Available: https://www.cdc.gov/mmwr/preview/mmwrhtml/mm5322a8.htm. Accessed 11 Nov 2019.
  5. 5. Danyluk MD, Jones TM, Abd SJ, Schlitt-Dittrich F, Jacobs M, Harris LJ. Prevalence and amounts of Salmonella found on raw California almonds. J Food Prot. 2007;70: 820–827. pmid:17477248
  6. 6. Bansal A, Jones TM, Abd SJ, Danyluk MD, Harris LJ. Most-probable-number determination of Salmonella levels in naturally contaminated raw almonds using two sample preparation methods. J Food Prot. 2010;73: 1986–1992. pmid:21219709
  7. 7. Santillana Farakos SM, Pouillot R, Johnson R, Spungen J, Son I, Anderson N, et al. A quantitative assessment of the risk of human salmonellosis arising from the consumption of almonds in the United States: The impact of preventive treatment levels. J Food Prot. 2017;80: 863–878. pmid:28414255
  8. 8. Parker CT, Huynh S, Quinones B, Harris LJ, Mandrell RE. Comparison of genotypes of Salmonella enterica serovar Enteritidis phage type 30 and 9c strains isolated during three outbreaks associated with raw almonds. Appl Environ Microbiol. 2010;76: 3723–3731. pmid:20363782
  9. 9. Brown E, Dessai U, Mcgarry S, Gerner-Smidt P. Use of whole-genome sequencing for food safety and public health in the United States. Foodborne Pathog Dis. 2019;16: 441–450. pmid:31194586
  10. 10. Stevens EL, Carleton HA, Beal J, Tillman GE, Lindsey RL, Lauer AC, et al. Use of whole genome sequencing by the Federal Interagency Collaboration for Genomics for Food and Feed Safety in the United States. J Food Prot. 2022;85: 755–772. pmid:35259246
  11. 11. Pightling AW, Pettengill JB, Luo Y, Baugher JD, Rand H, Strain E. Interpreting whole-genome sequence analyses of foodborne bacteria for regulatory applications and outbreak investigations. Front Microbiol. 2018;9: 1482. pmid:30042741
  12. 12. Davis S, Pettengill JB, Luo Y, Payne J, Shpuntoff A, Rand H, et al. CFSAN SNP Pipeline: an automated method for constructing SNP matrices from next-generation sequence data. PeerJ Comput Sci. 2015;1: e20.
  13. 13. Katz LS, Griswold T, Williams-Newkirk AJ, Wagner D, Petkau A, Sieffert C, et al. A comparative analysis of the Lyve-SET phylogenomics pipeline for genomic epidemiology of foodborne pathogens. Front Microbiol. 2017;8: 375. pmid:28348549
  14. 14. Wang YU, Pettengill JB, Pightling A, Timme R, Allard M, Strain E, et al. Genetic diversity of Salmonella and Listeria isolates from food facilities. J Food Prot. 2018;81: 2082–2089. pmid:30485763
  15. 15. Lambertini E, Danyluk MD, Schaffner DW, Winter CK, Harris LJ. Risk of salmonellosis from consumption of almonds in the North American market. Food Res Int. 2012;45: 1166–1174.
  16. 16. Emond-Rheault JG, Hamel J, Jeukens J, Freschi L, Kukavica-Ibrulj I, Boyle B, et al. The Salmonella enterica plasmidome as a reservoir of antibiotic resistance. Microorganisms. 2020;8: 1016. pmid:32650601
  17. 17. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30: 2114–2120. pmid:24695404
  18. 18. 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: 455–477. pmid:22506599
  19. 19. Zhang S, Yin Y, Jones MB, Zhang Z, Deatherage Kaiser BL, Dinsmore BA, et al. Salmonella serotype determination utilizing high-throughput genome sequencing data. J Clin Microbiol. 2015;53: 1685–92. pmid:25762776
  20. 20. Treangen TJ, Ondov BD, Koren S, Phillippy AM. The Harvest suite for rapid core-genome alignment and visualization of thousands of intraspecific microbial genomes. Genome Biol. 2014;15: 524. pmid:25410596
  21. 21. Price MN, Dehal PS, Arkin AP. FastTree 2 –Approximately maximum-likelihood trees for large alignments. PLoS One. 2010;5: e9490. pmid:20224823
  22. 22. Letunic I, Bork P. Interactive Tree Of Life (iTOL) v4: recent updates and new developments. Nucleic Acids Res. 2019;47: W256–W259. pmid:30931475
  23. 23. Kumar S, Stecher G, Tamura K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33: 1870–1874. pmid:27004904
  24. 24. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4: 406–425. pmid:3447015
  25. 25. Bortolaia V, Kaas RS, Ruppe E, Roberts MC, Schwarz S, Cattoir V, et al. ResFinder 4.0 for predictions of phenotypes from genotypes. J Antimicrob Chemother. 2020;75: 3491–3500. pmid:32780112
  26. 26. Florensa AF, Kaas RS, Clausen PTLC, Aytan-Aktug D, Aarestrup FM. ResFinder-an open online resource for identification of antimicrobial resistance genes in next-generation sequencing data and prediction of phenotypes from genotypes. 2022; pmid:35072601
  27. 27. Zankari E, Allesøe R, Joensen KG, Cavaco LM, Lund O, Aarestrup FM. PointFinder: A novel web tool for WGS-based detection of antimicrobial resistance associated with chromosomal point mutations in bacterial pathogens. J Antimicrob Chemother. 2017;72: 2764–2768. pmid:29091202
  28. 28. Liu B, Zheng D, Jin Q, Chen L, Yang J. VFDB 2019: a comparative pathogenomic platform with an interactive web interface. Nucleic Acids Res. 2019;47: D687–D692. pmid:30395255
  29. 29. Robertson J, Nash JHE. MOB-suite: software tools for clustering, reconstruction and typing of plasmids from draft assemblies. Microb Genomics. 2018;4. pmid:30052170
  30. 30. Robertson J, Bessonov K, Schonfeld J, Nash JHE. Universal whole-sequence-based plasmid typing and its utility to prediction of host range and epidemiological surveillance. Microb Genomics. 2020;6: 1–12. pmid:32969786
  31. 31. Robertson J, Yoshida C, Kruczkiewicz P, Nadon C, Nichani A, Taboada EN, et al. Comprehensive assessment of the quality of Salmonella whole genome sequence data available in public sequence databases using the Salmonella in silico Typing Resource (SISTR). Microb genomics. 2018;4. pmid:29338812
  32. 32. Turcotte MR, Smith JT, Li J, Zhang X, Wolfe KL, Gao F, et al. Genome characteristics of clinical Salmonella enterica population from a state public health laboratory, New Hampshire, USA, 2017–2020. BMC Genomics. 2022;23: 1–11. pmid:35870884
  33. 33. Ibrahim GM, Morin PM. Salmonella serotyping using whole genome sequencing. Front Microbiol. 2018;9: 1–8. pmid:30619114
  34. 34. Banerji S, Simon S, Tille A, Fruth A, Flieger A. Genome-based Salmonella serotyping as the new gold standard. Sci Rep. 2020;10: 1–10. pmid:32152449
  35. 35. Worley J, Meng J, Allard MW, Brown EW, Timme RE. Salmonella enterica phylogeny based on whole-genome sequencing reveals two new clades and novel patterns of horizontally acquired genetic elements. MBio. 2018;9. pmid:30482836
  36. 36. Chattaway MA, Dallman TJ, Larkin L, Nair S, McCormick J, Mikhail A, et al. The transformation of reference microbiology methods and surveillance for Salmonella with the use of whole genome sequencing in England and Wales. Front Public Heal. 2019;7: 317. pmid:31824904
  37. 37. Saltykova A, Wuyts V, Mattheus W, Bertrand S, Roosens NHC, Marchal K, et al. Comparison of SNP-based subtyping workflows for bacterial isolates using WGS data, applied to Salmonella enterica serotype Typhimurium and serotype 1,4,[5],12:i:-. Mossong J, editor. PLoS One. 2018;13: e0192504. pmid:29408896
  38. 38. Timme RE, Strain E, Baugher JD, Davis S, Gonzalez-Escalona N, Leon MS, et al. Phylogenomic pipeline validation for foodborne pathogen disease surveillance. J Clin Microbiol. 2019;57: e01816–18. pmid:30728194
  39. 39. Stasiewicz MJ, Oliver HF, Wiedmann M, den Bakker HC. Whole-genome sequencing allows for improved identification of persistent Listeria monocytogenes in food-associated environments. Appl Environ Microbiol. 2015;81: 6024–37. pmid:26116683
  40. 40. Register Federal. Almonds grown in California; outgoing quality control requirements. 7 CFR part 981. Fed Regist. 2007;72: 15021–15036. Available from: https://www.federalregister.gov/documents/2007/03/30/07-1557/almonds-grown-in-california-outgoing-quality-control-requirements
  41. 41. Haendiges J, Davidson GR, Pettengill JB, Reed E, Ramachandran P, Blessington T, et al. Genomic evidence of environmental and resident Salmonella Senftenberg and Montevideo contamination in the pistachio supply-chain. PLoS One. 2021;16: e0259471. pmid:34735518
  42. 42. Russo ET, Biggerstaff G, Hoekstra RM, Meyer S, Patel N, Miller B, et al. A recurrent, multistate outbreak of Salmonella serotype Agona infections associated with dry, unsweetened cereal consumption, United States, 2008. J Food Prot. 2013;76: 227–230. pmid:23433369
  43. 43. Harris LJ, Lieberman V, Mashiana RP, Atwill E, Yang M, Chandler JC, et al. Prevalence and amounts of Salmonella found on raw California inshell pistachios. J Food Prot. 2016;79: 1304–1315. pmid:27497117
  44. 44. Carattoli A. Plasmids and the spread of resistance. Int J Med Microbiol. 2013;303: 298–304. pmid:23499304
  45. 45. Ondov BD, Treangen TJ, Melsted P, Mallonee AB, Bergman NH, Koren S, et al. Mash: Fast genome and metagenome distance estimation using MinHash. Genome Biol. 2016;17: 1–14. pmid:27323842
  46. 46. Haendiges J, Keller S, Suehr Q, Anderson N, Reed E, Zheng J, et al. Complete genome sequences of five Salmonella enterica strains used in knoculation cocktails in low-moisture food storage studies. Microbiol Resour Announc. 2019;8: e01588–18.
  47. 47. Schonfeld J, Clark C, Robertson J, Arya G, Eagle SHC, Gurnik S, et al. Complete Genome Sequences for 36 Canadian Salmonella enterica serovar Typhimurium and I 1,4,[5],12:i:–isolates from clinical and animal sources. Microbiol Resour Announc. 2021;10: 9–11. doi:10.1128/mra.00734-20
  48. 48. Villa L, García-Fernández A, Fortini D, Carattoli A. Replicon sequence typing of IncF plasmids carrying virulence and resistance determinants. J Antimicrob Chemother. 2010;65: 2518–2529. pmid:20935300
  49. 49. McMillan EA, Jackson CR, Frye JG. Transferable Plasmids of Salmonella enterica associated with antibiotic resistance genes. Front Microbiol. 2020;11. pmid:33133037
  50. 50. Nguyen S V., Harhay DM, Bono JL, Smith TPL, Fields PI, Dinsmore BA, et al. Complete and closed genome sequences of 10 Salmonella enterica subsp. enterica serovar Anatum isolates from human and bovine sources. Genome Announc. 2016;4: 3–4. pmid:27257192
  51. 51. Caza M, Lépine F, Milot S, Dozois CM. Specific roles of the iroBCDEN genes in virulence of an avian pathogenic Escherichia coli O78 strain and in production of salmochelins. Infect Immun. 2008;76: 3539–3549. pmid:18541653
  52. 52. Li C, Pan D, Li M, Wang Y, Song L, Yu D, et al. Aerobactin-mediated iron acquisition enhances biofilm formation, oxidative stress resistance, and virulence of Yersinia pseudotuberculosis. Front Microbiol. 2021;12: 1–14. pmid:34335534
  53. 53. Fricke WF, McDermott PF, Mammel MK, Zhao S, Johnson TJ, Rasko DA, et al. Antimicrobial resistance-conferring plasmids with similarity to virulence plasmids from avian pathogenic Escherichia coli strains in Salmonella enterica serovar Kentucky isolates from poultry. Appl Environ Microbiol. 2009;75: 5963–5971. pmid:19648374
  54. 54. Tasmin R, Hasan NA, Grim CJ, Grant A, Choi SY, Alam MS, et al. Genotypic and phenotypic characterization of multidrug resistant Salmonella Typhimurium and Salmonella Kentucky strains recovered from chicken carcasses. Cloeckaert A, editor. PLoS One. 2017;12: e0176938. pmid:28481935
  55. 55. Neuert S, Nair S, Day MR, Doumith M, Ashton PM, Mellor KC, et al. Prediction of phenotypic antimicrobial resistance profiles from whole genome sequences of non-typhoidal Salmonella enterica. Front Microbiol. 2018;9: 592. pmid:29636749
  56. 56. Hurtado R, Barh D, Weimer BC, Viana MVC, Profeta R, Sousa TJ, et al. WGS-Based lineage and antimicrobial resistance pattern of Salmonella Typhimurium isolated during 2000–2017 in Peru. Antibiotics. 2022;11: 1170.
  57. 57. Tang B, Elbediwi M, Nambiar RB, Yang H, Lin J, Yue M. Genomic characterization of antimicrobial-resistant Salmonella enterica in duck, chicken, and pig farms and retail markets in Eastern China. Microbiol Spectr. 2022;10. pmid:36047803
  58. 58. Rehman MA, Yin X, Persaud-Lachhman MG, Diarra MS. First detection of a fosfomycin resistance gene,fosA7, in Salmonella enterica serovar Heidelberg isolated from broiler chickens. Antimicrob Agents Chemother. 2017;61. pmid:28533247
  59. 59. McDermott PF, Tyson GH, Kabera C, Chen Y, Li C, Folster JP, et al. Whole-genome sequencing for detecting antimicrobial resistance in nontyphoidal Salmonella. Antimicrob Agents Chemother. 2016;60: 5515–20. pmid:27381390
  60. 60. Cheng RA, Eade CR, Wiedmann M. Embracing Diversity: Differences in virulence mechanisms, disease severity, and host adaptations contribute to the success of nontyphoidal Salmonella as a foodborne pathogen. Front Microbiol. 2019;10: 1–20. pmid:31316476
  61. 61. Hurtado-Escobar GA, Grépinet O, Raymond P, Abed N, Velge P, Virlogeux-Payant I. H-NS is the major repressor of Salmonella Typhimurium Pef fimbriae expression. Virulence. 2019;10: 849–867. pmid:31661351
  62. 62. Libby SJ, Adams LG, Ficht TA, Allen C, Whitford HA, Buchmeier NA, et al. The spv genes on the Salmonella Dublin virulence plasmid are required for severe enteritis and systemic infection in the natural host. Infect Immun. 1997;65: 1786–1792.
  63. 63. Guiney DG, Fierer J. The role of the spv genes in Salmonella pathogenesis. Front Microbiol. 2011;2: 1–10. pmid:21716657
  64. 64. Krukonis ES, Thomson JJ. Complement evasion mechanisms of the systemic pathogens Yersiniae and Salmonellae. FEBS Lett. 2020;594: 2598–2620. pmid:32170725