ArticlePDF Available

Chicken Production and Human Clinical Escherichia coli Isolates Differ in Their Carriage of Antimicrobial Resistance and Virulence Factors

American Society for Microbiology
Applied and Environmental Microbiology
Authors:

Abstract and Figures

Pathogenic Escherichia coli causes disease in both humans and food-producing animals. E. coli pathogenesis is dependent on a repertoire of virulence factors and antimicrobial resistance genes. Food-borne outbreaks are highly associated with the consumption of undercooked and contaminated food products.
This content is subject to copyright. Terms and conditions apply.
Chicken Production and Human Clinical Escherichia coli Isolates
Differ in Their Carriage of Antimicrobial Resistance and
Virulence Factors
Reed Woyda,aAdelumola Oladeinde,bZaid Abdoa,c
a
Program of Cell and Molecular Biology, Colorado State University, Fort Collins, Colorado, USA
b
U.S. National Poultry Research Center, USDA-ARS, Athens, Georgia, USA
c
Department of Microbiology, Immunology, and Pathology, Colorado State University, Fort Collins, Colorado, USA
ABSTRACT Contamination of food animal products by Escherichia coli is a leading cause
of foodborne disease outbreaks, hospitalizations, and deaths in humans. Chicken is the
most consumed meat both in the United States and across the globe according to the
U.S. Department of Agriculture. Although E. coli is a ubiquitous commensal bacterium of
the guts of humans and animals, its ability to acquire antimicrobial resistance (AMR) genes
and virulence factors (VFs) can lead to the emergence of pathogenic strains that are resist-
ant to critically important antibiotics. Thus, it is important to identify the genetic factors
that contribute to the virulence and AMR of E. coli. In this study, we performed in-depth
genomic evaluation of AMR genes and VFs of E. coli genomes available through the
National Antimicrobial Resistance Monitoring System GenomeTrackr database. Our objec-
tive was to determine the genetic relatedness of chicken production isolates and human
clinical isolates. To achieve this aim, we rst developed a massively parallel analytical pipe-
line (Reads2Resistome) to accurately characterize the resistome of each E. coli genome,
including the AMR genes and VFs harbored. We used random forests and hierarchical clus-
tering to show that AMR genes and VFs are sufcient to classify isolates into different
pathogenic phylogroups and host origin. We found that the presence of key type III secre-
tion system and AMR genes differentiated human clinical isolates from chicken production
isolates. These results further improve our understanding of the interconnected role AMR
genes and VFs play in shaping the evolution of pathogenic E. coli strains.
IMPORTANCE Pathogenic Escherichia coli causes disease in both humans and food-pro-
ducing animals. E. coli pathogenesis is dependent on a repertoire of virulence factors
and antimicrobial resistance genes. Food-borne outbreaks are highly associated with the
consumption of undercooked and contaminated food products. This association high-
lights the need to understand the genetic factors that make E. coli virulent and patho-
genic in humans and poultry. This research shows that E. coli isolates originating from
human clinical settings and chicken production harbor different antimicrobial resistance
genes and virulence factors that can be used to classify them into phylogroups and
host origins. In addition, to aid in the repeatability and reproducibility of the results pre-
sented in this study, we have made a public repository of the Reads2Resistome pipeline
and have provided the accession numbers associated with the E. coli genomes analyzed.
KEYWORDS Escherichia coli, antimicrobial resistance, poultry production, random
forests, virulence factors
Escherichia coli are ubiquitous commensal bacteria in the gut of both humans and
food-producing animals and rarely cause disease but may acquire antimicrobial re-
sistance (AMR) genes and virulence factors (VFs) resulting in increased pathogenicity
(1). Pathogenic E. coli has consistently ranked in the top ve causative agents of
Editor Martha Vives, Unversidad de los Andes
Copyright © 2023 Woyda et al. This is an
open-access article distributed under the terms
of the Creative Commons Attribution 4.0
International license.
Address correspondence to Zaid Abdo,
zaid.abdo@colostate.edu, Adelumola
Oladeinde, ade.oladeinde@usda.gov, or Reed
Woyda, reed.woyda@colostate.edu.
The authors declare no conict of interest.
[This article was published on 18 January 2023
with a standard copyright line (© 2023
American Society for Microbiology. All Rights
Reserved.). The authors elected to pay for open
access for the article after publication,
necessitating replacement of the original
copyright line, and this change was made on 30
January 2023.]
Received 13 July 2022
Accepted 17 December 2022
Published 18 January 2023
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 1
ENVIRONMENTAL MICROBIOLOGY
disease outbreaks, outbreak-associated illnesses, and hospitalizations in the United
States (24) and is responsible for billions of dollars of annual health care associated
costs in the United States (5, 6). Pathogenic E. coli infections in humans may result in
acute to severe diarrhea or dysentery, urinary tract infections, and meningitis (79). E.
coli pathogenesis is dependent on the VF repertoire, which enables the bacterium to
evade host defenses, adhere to host surfaces, and successfully invadeand replicate
inhost tissues. For example, VFs such as toxins, iron-acquisition systems, and m-
briae play integral roles in the pathogenicity of extraintestinal E. coli strains that cause
urinary tract infections in humans (10). These uropathogenic E. coli strains are able to
colonize human mucosal surfaces due to surface adhesion VFs, including the P, F, and
type 1 mbriae encoded by pap,sfa, and mgenes (11).
Outbreaks in human populations caused by pathogenic E. coli are attributed to the con-
sumption of undercooked and contaminated foods, including meats and vegetables (12).
E. coli infections in food-producing animals usually results in diarrhea but can include acute
fatal septicemia, airsacculitis, pericarditis, and perihepatitis (13). In poultry, avian-patho-
genic E. coli infections, termed avian colibacillosis, can result in economic loss due to the
cost of treatment, as well as from culling of ocks (14). Colibacillosis is a leading cause of
morbidity and mortality in poultry, as noted by decreases in the feed conversion ratio, egg
production, and hatching rates (1517). The severity of the disease depends on the VF rep-
ertoire of the strain, including genes encoding iron acquisition and transport systems (18).
Extraintestinal pathogenic E. coli strains that cause urinary tract infections, neonatal
meningitis, and sepsis have VFs similar to those of E. coli isolates from meat and animal
sources (19). Likewise, some avian-pathogenic E. coli and extraintestinal pathogenic E.
coli isolates have been reported to share similar VFs and belong to the same multilocus
sequence type (MLST) and phylogroup (2023). Specically, a microarray-based study
of E. coli isolates from both human and animal sources in Denmark identied 66 to 87
genes, including both virulence factors and antimicrobial resistance genes, which were
present both in human urinary tract infection isolates and in isolates obtained from
poultry and pig products (24). Genes such as the iron acquisition genes iutA and iroN,
as well as mbrial papA encoding the Pap mbrial major subunit and adhesion genes
(e.g., papGII encoding Pap adhesion), were detected in both urinary tract infection and
poultry-related isolates (24). In vivo and in vitro experimental studies have shown that
extraintestinal pathogenic E. coli recovered from humans can cause disease in avian
hosts and, similarly, avian-pathogenic E. coli isolates recovered from avian host can
cause disease in mammalian models (2529).
Infections with pathogenic E. coli are commonly treated with antibiotics; however,
increasing levels of AMR impose difculties in selecting effective treatment options
(30). The spread and increased prevalence of AMR has been linked to both the overuse
and the misuse of antibiotics in human clinical settings, as well as in food animal pro-
duction (31). Consumer opinion has led to a reduction in antimicrobial use in food ani-
mal production to mitigate the spread of AMR. However, recent studies have indicated
that AMR can still persist in food animal production even after the removal or stoppage
of antibiotics (32, 33). The comobilization and coacquisition of AMR genes and VFs
through horizontal gene transfer can result in highly pathogenic E. coli strains. Recent
studies have demonstrated a correlation and close genetic linkage of AMR genes and
VFs in pathogenic bacterial strains (34, 35). Pan et al. (35) performed a comprehensive
analysis of over 9,000 bacterial genomes from multiple species and hosts and observed
the coexistence of AMR genes and VFs from human-associated pathogens. Therefore,
the aim of this study was to evaluate and compare the current distribution of AMR
genes and VFs present in E. coli isolates from chicken production and human clinical
settings in the United States.
To do so, we determined the AMR genes and VFs present in nearly 800 E. coli genomes
collected from chicken production and the human clinical settings. Data for all isolates
were obtained from the National Antimicrobial Resistance Monitoring System (NARMS).
We used the most recent World Health Organization (WHO) classication of antimicrobials
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 2
important to human health to determine the extent to which resistance to these antimicro-
bials are prevalent in chicken production and human clinical settings. We hypothesized
that resistance to antimicrobials of critical and high importance to human health would be
present in both human- and chicken-derived isolates. In addition, we hypothesized that E.
coli isolates from human clinical settings and chicken production will differ in their carriage
of AMR genes and VFs. We found that E. coli isolates obtained from chicken production
and human clinical settings harbored AMR genes predicted to confer resistance to many
antimicrobials classied as highly and critically important to human health by the WHO.
Furthermore, we were able to classify all E. coli isolates into their respective phylogroups
and host origins using random forest classication and hierarchical clustering of the AMR
genes and VFs found in their genomes.
RESULTS
E. coli isolate selection and genome assembly. To evaluate and compare the cur-
rent distribution of AMR genes and VFs present in E. coli isolates from chicken produc-
tion and the human clinical settings in the United States, we took advantage of the
U.S. National Antimicrobial Resistance Monitoring System (NARMS) GenomeTrackr
database (available through the NCBI [https://www.ncbi.nlm.nih.gov/pathogens]). We
chose NARMS because it monitors enteric bacteria and pathogens across all states to
determine their resistance to antimicrobials used in veterinary and human medicine.
NARMS is a collaboration between agencies within the U.S. Department of Health and
Human Services (HHS; Food and Drug Administration [FDA] and Centers for Disease
Control and Prevention [CDC]) and the U.S. Department of Agriculture (USDA; Food
Safety and Inspection Service, Animal and Plant Health Inspection Service, and the
Agricultural Research Service). Each USDA agency taking part in NARMS tests bacterial
samples at various stages of food animal production. The NARMS database holds bac-
terial genome sequences that originate from 15 distinct human, animal, and food sour-
ces and are categorized into four NARMS samples sources: (i) human clinical isolates;
(ii) food animal production isolates from cecal samples at slaughter; (iii) samples rou-
tinely collected at inspected establishments as part of FSIS verication testing; and (iv)
raw meats (chicken, ground turkey, ground beef, and pork chops) collected at retail
outlets in 20 states (23 sites). We selected 837 chicken production isolates and 874
human clinical isolates, with collection dates from 1 January 2018 until 30 March 2020,
using species,”“host,and locationsearch lters (see Materials and Methods for a
detailed description). E. coli isolate data were either available as Illumina or Oxford
Nanopore FASTQ reads or as FASTA assemblies.
A total of 130 FASTA assemblies were obtained for human clinical isolates, whereas all
chicken production isolates and remaining human clinical isolates had available FASTQ
reads. The FASTA assemblies were obtained through the isolates associated BioProject,
and quality control was done using QUAST (36). The FASTQ reads for each isolate were rst
normalized to 30coverage using Trinitysinsilico_read_normalization.pl script (https://
github.com/trinityrnaseq/trinityrnaseq/wiki). After quality control, the average genome
coverage was 35.8and 31.7for human clinical and chicken production isolates, respec-
tively (Fig. 1A). Totals of 385 chicken production isolates and 535 human clinical isolates
were excluded from further analysis because they did not meet our preset quality control
metrics based on genome coverage $20and assembled genome quality (N
50
$10,000
bp; Fig. 1B). The remaining E. coli isolates from chicken production and human clinical set-
tings were conrmed to be E. coli through MLST (see Table S1, https://github.com/
tseemann/mlst). This resulted in our study population consisting of 452 (57%) chicken pro-
duction isolates and 339 (43%) human clinical isolates (Fig. 2A and B). Isolates originated
from a diverse set of sources (see Table S1 in the supplemental material) and locations
within the United States (Fig. 2C).
Reads2Resistome enables high-throughput genome assembly and resistome
characterization. We developed Reads2Resistome (R2R) (37) to streamline the quality
control, assembly, annotation, and resistome characterization of the chicken production
and human clinical genomes (see the supplemental material for detailed description of
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 3
Reads2Resistome [Fig. S1]). Utilizing a singularity (38) container, R2R maintains software
and database versions resulting in reproducible results that can be replicated using the
same input data and command-line options. We rst assessed the accuracy and perform-
ance of R2R using short (Illumina) and long FASTQ reads (PacBio and Oxford Nanopore) of
Salmonella enterica serovar Heidelberg and E. coli isolates recovered from the ceca of
broiler chickens (see Table S2). Short and long reads-only assemblies resulted in the short-
estrun-timewithanaverageof6minpersample,whilehybridassemblies(shortandlong
reads were combined) required ;1 h per sample (see Table S3). Genomes assembled with
short reads only or a hybrid approach had better resistome characterization and gene
annotation than genomes assembled using long reads only (see Tables S4 and S5). For this
study, all E. coli isolates were assembled from short Illumina reads except for a single isolate
with long-read data (SRR11174251).
AMR gene distribution not explained by isolation year or location. We con-
ducted principal-component analysis (PCA) to determine whether the presence or absence
of identied AMR genes corresponded to the state origin of the isolates. Overlapping 95%
condence intervals were observed for all isolates in all states (see Fig. S1). We also con-
ducted PCA on the presence/absence matrix of these AMR genes to assess impact of time
of sampling. Separation of isolates based on their associated isolation year was also not
observed (see Fig. S2). Accordingly, year and location will not be included in further analysis.
Determining the phylogeny of E. coli isolate phylogroups. To determine whether
there is an evolutionary based separation between chicken production and human
clinical isolates, we performed a whole-genome alignment to the E. coli K12-MG1655
reference genome (accession U00096). Single-nucleotide polymorphisms (SNPs) with
FIG 2 Graphical description of human clinical and poultry production isolates. (A and B) FASTQ reads or FASTA assemblies for each
E. coli isolate was downloaded from the NARMS GenomeTrackr for chicken production and human clinical isolates. Isolates were
selected from submission dates of 1 January 2018 to 30 March 2020 using the ltering criteria described in Materials and Methods.
After quality control and ltering, 452 chicken production isolates (57%) and 399 human clinical isolates (43%) remained. (C) Isolate
geographiclocationsacrosstheUnitedStates.MapsweregeneratedusingtheFilled Mapchart type in Microsoft Excel from Microsoft
Ofce 365 (Microsoft Corporation).
FIG 1 Average coverage, postnormalization, and average N
50
, postassembly, of human clinical and
poultry production isolates. (A) Isolate FASTQ reads were downloaded from the NCBI and normalized
to 30using Trinitysinsilico_read_normalization.pl script. (B) Isolates with available FASTQ reads
were assembled and annotated using Reads2Resistome, including N
50
calculation by QUAST. FASTA
assemblies were annotated using the additional script provided by Reads2Resistome for postassembly
annotation. A diamond shape indicates a mean value, and the horizontal lines within a box indicate
the mean. The upper and lower sides of the box correspond to the 25th and 75th percentiles. Plots
were generated in R v4.0.4 using ggplot2 v3.3.3.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 4
respect to the reference genome were used to generate a phylogenetic tree. In addi-
tion, we used ClermonTyping (39) to classify the isolates into phylogroups. E. coli iso-
lates within the phylogenetic tree clustered into single clades and chicken production
isolates were less divergent from the E. coli K-12 reference genome than human clinical
isolates (Fig. 3). E. coli isolates were phylotyped into one of the phylogroups A, B1, B2,
C, clade I, D, E, E or clade I,F, and G (Table 1). The majority of the E. coli genomes in
our data set belonged to phylogroups A, B1, and B2. Phylogroups B1 and B2 were the
most represented groups, with each accounting for ;23% of the total E. coli isolates.
Phylogroup determination for one isolate was unsuccessful (identied as unknown),
but the isolate clustered with phylogroup A isolates on the SNP-based tree.
Phylogroup classication is useful for the rapid and easy identication of potentially
FIG 3 Whole-genome SNP-based maximum-likelihood phylogenetic tree. SNPs found between the E. coli K12-MG1655
(accession U00096) reference genome, and each of the 791 E. coli isolates was used for alignment using Snippy v4.6.0
(Snippy 2018). The resulting whole-genome SNP alignment was used to construct maximum-likelihood phylogenetic
trees under a GTR1GAMMA model using RAxML with 100 bootstraps (RAxMLHPC-PTHREADS, v8.2.12). The strength of
nodal support is indicated by the branch color: red (100), orange (50.to #75), dark orange (25.to #50), and
green (0.to #25). (Ring) Phylogroups identied using ClermonTyping v1.4.0. Phylogroups are labeled by color: A
(green), B1 (orange), B2 (magenta), C (brown), clade I (aquamarine), D (blue), E (gray), E or clade I (maroon), F (gold), G
(dark gold), and unknown (dark green).
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 5
virulent and AMR isolates (39). Based on the Clermont quadruplex in silico method, our
studyisolatesrepresentedallsevenmainphylogroups(A,B1,B2,C,D,E,andF),cryptic
clade I, and the newly identied phylogroup G. One isolate was classied as unknown. We
found signicant differences in the proportion of times phylogroups A, B1, B2, E, and G
were seen with respect to their host origin. Phylogroups A, B1, and G were found at higher
frequencies in chicken production isolates, while B2 and E were found at higher frequencies
in human clinical isolates (Table 1). Eighty-one percent of chicken production isolates
belonged to phylogroup B1, whereas 85% of human clinical isolates were classied as B2.
The most divergent subclade within B2 consisted of closely related human clinical isolates,
many of which originated from the same BioProject (PRJNA489090) which included many
isolates from the human infant gut. Although the B2 phylogroup was dominated by human
clinical isolates, there were a few chicken isolates nested within the B2 clades. Phylogroup E
was dominated by closely related human clinical isolates (81%) and was the largest cluster
of human clinical isolates within the phylogenetic tree. Isolates identied as belonging to
the Escherichia cryptic clade I phylogroup (NCBI genome accession numbers SRR9984913,
SRR10687720,SRR9852843,SRR9875260,SRR9875260,SRR10267809,SRR10687983,and
SRR9852805)(40) clustered together within the phylogenetic tree and were all of chicken or-
igin. No human clinical E. coli isolates were identied as belonging to clade I, even though
clade I has been reported to have 2 to 3% prevalence in human E. coli isolates (4143).
Phylogroup G was dominated by chicken production isolates (95%).
The analysis presented above shows that in most cases the phylogroup classica-
tion supported the phylogenetic distribution within the SNP-based tree; however, we
found several instances where human clinical and chicken production isolates were
closely related. Likewise, we found cases where the SNP-based classication contra-
dicted the phylogroup classication. For instance, human clinical strains mainly belong
to phylogroup A (4143); however, we identied 85.58% of our chicken production iso-
lates belong to this phylogroup. Taken together, our isolate collection serves as a good
representation and distribution of the currently described E. coli phylogroups, with a
few exceptions.
Chicken production and human clinical isolates harbor a diverse set of AMR
genes and mutations. We used the ResFinder (44) to identify acquired AMR genes in
each E. coli genome. In addition, we used R2R and the Comprehensive Antimicrobial
Resistance Database (CARD) via the Resistance Gene Identier(RGI)(45)(seeMaterialsand
Methods for details) to identify all AMR genes, including those encoded on the bacterial
chromosome and gene mutations that can result in resistant genotypes. We found 42
acquired AMR genes predicted to confer resistance to 21 drug classes (Table 2). Fifty-seven
percent of the isolates carried at least 1 acquired AMR gene, while one human clinical iso-
late carried 21 acquired AMR genes. Chicken production isolates were more likely to harbor
acquired AMR genes (65.7%) than human clinical isolates (46.3%) (adjusted P0). Chicken
TABLE 1 Distribution of phylogroups in chicken production and human clinical isolates
a
Phylogroup
No. of isolates (% total)
Total (n= 791) Chicken production (n= 452) Human clinical (n= 339)
A 132 (16.62) 109 (85.58)*23 (17.42)*
B1 183 (23.05) 148 (80.87)*35 (19.13)*
B2 184 (23.17) 28 (15.21)*156 (84.78)*
C 23 (2.90) 18 (78.26) 5 (21.74)
Clade I 8 (1.01) 8 (100) 0 (0)
D 103 (12.97) 56 (54.37) 47 (45.63)
E 75 (9.45) 14 (18.67)*61 (81.33)*
E or clade I 4 (0.05) 4 (100) 0 (0)
F 26 (3.27) 15 (57.69) 11 (42.31)
G 55 (6.93) 52 (94.55)*3 (5.45)*
Unknown 1 (0.13) 0 (0) 1 (100)
a
The phylogroup for each isolatewas determined using ClermonTyping v1.4.0 with default settings for each isolate
genome assembly. Phylogroups that are signicantly associated with a specic host are indicated with an asterisck
(*), as determined by maximum-likelihood ratio testing with a Benjamini-Hochberg adjusted Pvalue of ,0.05.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 6
production E. coli isolates carried more acquired AMR genes per isolate (range, 0 to 8;
mean, 1.97 per isolate) compared to human clinical isolates (range, 0 to 21; mean, 1.87 per
isolate). Fosfomycin (fosA7) and peptide antibiotics (mcr-9)weresignicantly higher in
chicken production isolates, while aminoglycosides [aac(69)-Ib-cr]anduoroquinolones
(qrnB19)weresignicantly higher in human clinical isolates (Fig. 4A; see also Table S6).
Using the CARD database we found 208 AMR genes, including all but three genes [aac
(39)-IIa,sitABCD,andformA] identied by ResFinder (see Table S1). A majority of AMR genes
were shared between human clinical isolates and chicken production isolates (Fig. 5A). The
AMR genes were predicted to confer resistance to 31 drug classes, including cephalosporins,
uoroquinolones, penams, and tetracyclines (Fig. 4B). We found AMR genes for 31 drug
classes in human clinical isolates, while 28 of the 31 drug classes were found in chicken pro-
duction isolates. AMR genes absent from chicken production isolates were predicted to con-
fer resistance to streptogramin (ermB and msrA), oxazolidinone (msrA), and pleuromutilin
(msrA and taeA). We found signicant differences in antimicrobial drug classes between
chicken production and human clinical isolates for 14 of the 31 drug classes (adjusted
P,0.05, Fig. 4B; see also Table S7). Genes predicted to confer resistance to diaminopyrimi-
dine, phenicol, sulfonamide, penem, aminoglycoside, uoroquinolone, macrolide, and pep-
tide antibiotics were more prevalent within human clinical isolates (adjusted P0, Fig. 4B).
Genes predicted to confer cephamycin resistance a highly important antimicrobial group
for human medicine (46) were more prevalent in human clinical isolates (adjusted P0,
Fig. 4B) compared to chicken isolates. Genes predicted to confer resistance to elfamycin and
nucleoside antibiotics were more prevalent within the chicken production isolates com-
pared to human isolates (adjusted P0, Fig. 4B). The frequency of genes predicted to con-
fer resistance to tetracycline was similar (adjusted P= 0.27, Fig. 4B) between host origins.
It is important to mention that the majority of AMR genes identied by CARD are
likely chromosomally encoded and may not be mobilizable. Many of the AMR genes
harbored on the chromosome confer levels of resistance that are not medically rele-
vant but may confer tness advantages to these isolates in their native environments
(47, 48). In contrast, most of the AMR genes identied by ResFinder are more likely to
be mobilized within bacterial populations, via plasmids or other mobile genetic ele-
ments (MGEs), are therefore more likely to spread between/within E. coli strains origi-
nating from human and food animal populations. We hypothesized that a majority of
TABLE 2 Acquired antimicrobial resistance genes identied within chicken production and human clinical isolate FASTA assemblies using
ResFinder
a
Gene Drug class Gene Drug class
aac(3)-IIa Aminoglycoside antibiotic dfrA1 Diaminopyrimidine antibiotic
aac(3)-IId Aminoglycoside antibiotic dfrA12 Diaminopyrimidine antibiotic
aac(3)-IV Aminoglycoside antibiotic dfrA14 Diaminopyrimidine antibiotic
aac(69)-Ib-cr Fluoroquinolone antibiotic, Aminoglycoside antibiotic dfrA17 Diaminopyrimidine antibiotic
aadA1 Aminoglycoside antibiotic erm(B) Lincosamide antibiotic, macrolide antibiotic, streptogramin antibiotic
aadA2 Aminoglycoside antibiotic oR Phenicol antibiotic
aadA5 Aminoglycoside antibiotic formA Disinfecting agents
aph(30)-Ib Aminoglycoside antibiotic fosA7 Fosfomycin
aph(39)-Ia Aminoglycoside antibiotic mcr-9 Peptide antibiotic
aph(39)-IIa Aminoglycoside antibiotic mdf(A) Rhodamine, tetracycline antibiotic, benzalkonium chloride
aph(39)-III Aminoglycoside antibiotic mph(A) Macrolide antibiotic
aph(4)-Ia Aminoglycoside antibiotic qacE Disinfecting agents and intercalating dyes
aph(6)-Ic Aminoglycoside antibiotic qnrB19 Fluoroquinolone antibiotic
aph(6)-Id Aminoglycoside antibiotic rmtB Aminoglycoside antibiotic
bla
CMY-2
Penam, carbapenem, cephalosporin, cephamycin sitABCD Disinfecting agents
bla
CTX-M-14
Cephalosporin sul1 Sulfonamide antibiotic
bla
CTX-M-15
Penam, cephalosporin sul2 Sulfonamide antibiotic
bla
OXA-1
Penam, carbapenem, cephalosporin sul3 Sulfonamide antibiotic
bla
OXA-244
Penam, carbapenem, cephalosporin tet(A) Tetracycline antibiotic
bla
TEM-1B
Monobactam, penam, penem, cephalosporin tet(B) Tetracycline antibiotic
catA1 Phenicol antibiotic tet(C) Tetracycline antibiotic
a
ResFinder output hits for each gene were ltered ($80% identity to database reference query). The corresponding resistance-conferring drug class for each gene was
identied using the Comprehensive Antibiotic Resistance Database (CARD).
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 7
the AMR genes identied by ResFinder were located either on plasmid or chromo-
somal contigs and not located in close proximity to other MGEs such as genomic
islands, insertion sequences, gene cassettes, etc. This hypothesis was proposed due to
all but three ResFinder identied AMR genes being also identied by CARD which are
majorly present on chromosomal contigs. To test this hypothesis, we submitted four
isolates to VRProle2 (49), an online bacterial mobile element detection pipeline capa-
ble of predicting AMR gene presence on plasmid or chromosomal contigs or those in
close proximity to other MGEs. VRProle2 results indicated AMR genes found by
ResFinder were located on plasmid contigs, as was the case for aph(39)-III. No AMR
genes were identied in close proximity to any other MGE types (see Table S8).
Overall, we observed a high prevalence of genes that can confer resistance to anti-
microbials considered highly and critically important to human medicine in both
chicken production and human clinical E. coli isolates. The prevalence of these genes
in both human and chicken E. coli isolates indicates that these populations may serve
as a possible reservoir of antimicrobial resistance.
The presence of type III secretion system genes classied the E. coli isolates
into two separate clusters. To determine whether AMR genes and VFs (Fig. 5B) could
separate the human clinical and chicken production isolates into distinct groups, we
employed hierarchical clustering. Hierarchical clustering was based on the ltered pres-
ence/absence table of AMR genes and VFs (Fig. 6; see also Table S1) and resulted in two
FIG 4 Average proportions of identied antimicrobial resistance genes in human clinical and poultry production isolates. (A) Acquired antimicrobial
resistance genes identied within chicken production and human clinical isolate FASTA assemblies using ResFinder. ResFinder output hits for each gene
were ltered ($80% identity to database reference query). (B) Antimicrobial resistance drug classes identied within chicken production and human clinical
isolate FASTA assemblies using Resistance Gene Identier (RGI) and Abricate employed via Reads2Resistome. AMR database hits were ltered ($95%
identity to database reference query), and the highest hit from each database was retained. Genes conferring resistance to drug classes were enumerated
for each isolate, and a proportion was calculated using the total number of genes in the study population conferring resistance to a given drug class. Drug
classes signicantly differing between chicken production isolates and human clinical isolates were determined by Wilcoxon rank sum test, and an asterisk
(*) indicates a Benjamini-Hochberg adjusted Pvalue of ,0.05.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 8
distinct clusters (A and B) consisting of 661 and 130 isolates, respectively. These two clus-
ters were signicantly different from one another (adjusted P=P0.05) compared by
isolate host origin. Cluster A comprised mostly of chicken production isolates (407 of 664
[61%]) while cluster B composed of human clinical isolates (85 of 130 [65%]).
To determine the genes driving the separation of these isolates into the two distinct
clusters, we performed random forest classication on the ltered AMR gene and VF ta-
ble (see Table S1). This reclassication resulted in 100% correct classication of all iso-
lates to the two clusters (A and B) (see Table S9). Type 3 secretion system (T3SS) genes
play a key role in the virulence of many Gram-negative bacterial pathogens (49) and
were prevalent in all isolates from cluster B and completely absent from cluster A iso-
lates. Twenty-ve percent of all human clinical isolates (n= 339) carried genes for the
T3SS operon (50), whereas only 9.29% of all chicken production isolates (n= 352) har-
bored the T3SS operon. Along with genes from the T3SS operon, the intimin gene
(eae) that is required for intimate adherence and virulence in both humans and animals
was a high contributor to the mean decreasing accuracy and was present in 100% of
cluster B isolates but completely absent from cluster A isolates (Fig. 7A). Reducing the
input AMR and virulence factors to only include the T3SS outer ring protein (escD) and
the T3SS secretin (escC), two genes that encode oligomerizing proteins for the T3SS
and that are essential for T3SS functionality (49), resulted in 100% correct classication
of all isolates to either cluster A or B. Our results suggest that cluster B isolates that har-
bor T3SS may have the potential to deliver effector proteins that promote bacterial col-
onization, replication, and transmission in human host cell cytoplasm, while cluster A
isolates may lack this capability. Nevertheless, we found both clusters contained iso-
lates from human clinical and chicken production settings.
AMR genes and virulence factors classied E. coli isolates by their host origin.
To assess which AMR genes and VFs differentiate isolates based on their host origin we
performed Random Forests classication (Fig. 7B). This classication resulted in 1.1% (5
isolates) error in classifying chicken isolates and a 5.6% (19 isolates) error in classifying
human clinical isolates (see Table S10). All misclassied isolates belonged to the
FIG 5 Venn diagrams depicting the percentage of antimicrobial resistance genes identied via
Reads2Resistome (RGI) and The Comprehensive Antibiotic Resistance Database and ResFinder (A) and
virulence factors identied via Reads2Resistome (Abricate) and the Virulence Factor Database (B).
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 9
hierarchical cluster A (see Table S11) described earlier. The E. coli topoisomerase IV
(parC) gene with the S80I mutation, was the highest contributor to the mean deceas-
ing accuracy. The parC gene plays a critical role in DNA replication and confers reduced
susceptibility to uoroquinolones if the relevant parC and gyrase A (gyrA) mutations
are present together (47, 5154). A total of 185 isolates (1 chicken production and 184
human clinical) had parC mutations, and 215 isolates (25 chicken production and 190
human clinical) had gyrA mutations. Four human clinical isolates had parC mutations
but no gyrA mutation, while 34 isolates (24 chicken production and 10 human clinical)
had gyrA mutations but no parC mutation. One E. coli isolate, misclassied as a human
isolate, harbored both parC (S80I) and gyrA (S83L, S84L, and D87Y) mutations (see
Table S12). The remaining 572 isolates did not have a parC or a gyrA mutation. In
FIG 6 Heatmap of AMR genes and virulence factors across all isolates (see Table S1). Heatmap was generated
in R v4.0.4 with pheatmap v1.0.12 (clustering_method = average[UPGMA], clustering_distance_cols =
binary,clustering_distance_rows = Euclidean) using the ltered AMR gene and virulence factor table.
Isolates from the phylogroups unknownand E or clade Iare highlighted.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 10
summary, parC and gyrA mutations that confer resistance to uoroquinolones were
present in many of the human clinical isolates, and it was a major factor that separated
the E. coli isolates into their respective host origins.
The iron acquisition genes (iroN,iroD,andiroE) were high contributors to the mean
decreasing accuracy of random forest classication. The catecholate siderophore uptake
system (iroBCDEN) plays a critical role in virulence since iron is required for many cellular
processes, but it is limited in host sites of extraintestinal infections (55). This gene cluster
has been reported to be located in a pathogenicity island on the chromosome and has
been found on plasmids (33, 56). However, this gene cluster was absent from all misclassi-
ed chicken isolates but was present in 56 and 7% of chicken production isolates and
human clinical isolates, respectively (see Table S1). The enterotoxin TieB protein (senB)was
the second most important factor and was absent from all chicken production isolates
regardless of their random forest classication but present in 28% of human clinical iso-
lates. Only one misclassied human clinical isolate harbored senB. These data suggest that
the human and chicken isolates in our data set carry distinct AMR determinants and VFs
which differentiate them from one another.
AMR genes and VF content can classify E. coli isolates by phylogroup. Next, we
hypothesized that if the AMR genes and VF content is specic to each phylogroup, we
should be able to classify the E. coli isolates correctly into their respective phylogroups
based on their AMR and VF proles. To evaluate this, we used random forests to clas-
sify the isolates into their identied phylogroups. Classication resulted in the follow-
ing error rates: phylogroup A (11%), B1 (3%), B2 (0%), C (48%), clade I (0%), D (0%), E
(5%), E or clade I(100%), F (4%), G (2%), and unknown (100%) (see Table S13). Unlike
FIG 7 Mean decreasing accuracy (MDA) plot of variables in random forest classication models. MDA represents how much accuracy
the model losses by exclusion of each variable (AMR gene or virulence factor). Random forest classication was performed on AMR
gene and virulence factor table (see Materials and Methods for details). (A) Resulting MDA plot after classication of isolates into
hierarchical clustering identied clusters A and B. Genes highlighted in red, on their own, result in 100% classication of isolates into
hierarchical clustering identied clusters A and B. (B) Resulting MDA plot after classication of isolates into chicken production
isolates and human clinical isolates. (C) Resulting MDA plot after classication of isolates into ClermonTyping identied phylogroups.
Random forest classication in R was performed using the randomForest v4.6-14. Asterisks (*) indicate genes with mutations: parC
(S80I), gyrA(S83L) (n= 30), gyrA(S84L) (n= 2), gyrA(D87Y) (n= 275), UtpT(E350Q), CyaA(S352T), and GlpT(E448K) (see Table S14).
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 11
host origin classication, reducing the input table for phylogroup classication resulted
in an increase in the misclassication error. Interestingly, the top 30 factors, needed to
classify isolates into their respective phylogroups, consisted of 8 AMR genes and 22
VFs (Fig. 7C), suggesting that VFs have a greater inuence than AMR genes for phy-
logroup classication.
Of 11 VF functions found using the virulence factor database (see Table S14), 6 were
present in all phylogroups, and the remaining functions were either entirely absent from
individual phylogroups or present in some (see Table S15). Interestingly, virulence func-
tions relating to adherence, secretion systems, and iron uptake were present in all phy-
logroups, but functions relating to immune evasion were present only in phylogroups A,
B1, B2, D, and E. (see Table S15). Thus, both chicken production and human clinical isolates
have the potential to serve as VF reservoirs. Albeit E. coli isolates of phylogroups C, Eor
clade I,cladeI,GandF,lackedmanyoftheVFfunctionscomparedtoA,B1,B2,D,andE.
We found that an isolatesAMRandVFprole was sufcient, in most cases, to deter-
mine its phylogroup. In addition, misclassied isolates were supported by their AMR and
VF similarity as seen using the UPGMA (unweighted pair-group method with arithmetic
averages) clustering approach (see Materials and Methods for details; Fig. 6). For instance,
the E. coli isolate that was identied by ClermonTyping as unknown was classied into phy-
logroup A based on its AMR gene and VF prole similarity with other phylogroup A, B1,
and C isolates (Fig. 6). The misclassication of EorcladeIisolates (NCBI accession num-
bers SRR10805089,SRR10579993,andSRR9262887) as phylogroups D and E, respectively,
were also supported by clustering in close proximity to other phylogroup D and E isolates
(Fig. 6). An additional EorcladeIisolate (NCBI accession number SRR9984544)wasmis-
classied as D but contained a similar AMR and VF prole as other phylogroup D isolates.
The misclassication was due to the absence of iron uptake genes shuS,shuA,andshuX,as
well as the non-locus of enterocyte effacement (LEE)-encoded T3SS genes espL4 and espY1.
Although the AMR and VFs harbored by an E. coli isolate provided sufcient signal for cor-
rect classication of phylogroups, it could not correctly classify all phylogroups identied
by ClermonTyping. However, the misclassication of unknowngroups and E or clade I
isolates was explained and supported by their clustering based on the similarity of their
AMR gene and VF proles (Fig. 6). Similarly, the location of the unknownisolate within
the phylogenetic tree supports the random forests classication (Fig. 3), i.e., clustered with
phylogroup A isolates.
DISCUSSION
In this study, we utilized E. coli isolates obtained from the NARMS surveillance pro-
gram to study differences in phylotype, AMR, and VFs between chicken production
and human clinical isolates. To do this, we rst developed a massively parallel pipeline
(Reads2Resistome) to streamline our genomic analysis and to ensure that our results
are repeatable and reproducible. We found that E. coli can serve as a reservoir of AMR
genes and VFs which supports previously reported ndings (5759). We determined
that a majority of AMR genes and VFs identied were shared (65 and 84%, respectively)
between human clinical isolates and chicken production isolates. Although there is a
possibility of horizontal gene transfer in the environment, the E. coli isolates from this
study were mostly host-specic, suggesting that strains from chicken production rarely
colonize humans. We identied AMR genes that have been acquired, as well as those
encoded on the chromosome that can acquire mutations that will result in reduced
susceptibility to antibiotics. We conrmed the presence of 42 acquired AMR genes con-
ferring resistance to 21 drug classes (Table 2, Fig. 4A, and Fig. 5A). The drug classes
identied include those considered critically and highly important to human medicine.
For example, genes conferring resistance to fosfomycin, considered a high priority and
critically important antimicrobial by the WHO (46), were signicantly more prevalent
within the chicken production isolates. Utilizing the CARD database, a collection of all
AMR genes, including those with mutations known to confer a level of resistance to
known antimicrobials, we identied genes conferring resistance to 31 antibiotic drug
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 12
classes in our human clinical isolates, 28 of which were observed in the chicken pro-
duction isolates (Fig. 4B). In addition, genes conferring resistance to these 31 drug
classes were observed across a majority of identied phylogroups.
We determined that chicken production isolates carried a higher proportion of acquired
AMR genes than human clinical isolates. Acquired AMR genes are those which have been
obtained by a strain via mobile genetic elements such as plasmids, integrative conjugative
elements, bacteriophages, insertion sequences, and gene capture elements such as a site-
specic recombination system (59). Acquired AMR genes are important in the context of
animal-to-human transmission since they are more readily shared between bacteria and
providearoutefordisseminationofAMRgenes.Thehighproportionofbenzalkonium
chloride is not surprising due to its high usage, as a disinfectant, in the food and medical
industries (60). The gene conferring resistance to benzalkonium chloride, mdfA,amultidrug
efux system, also confers resistance to tetracycline and rhodamine. When comparing the
presence of AMR genes from CARD, which includes genes with mutations resulting in a re-
sistance genotype to acquired resistance genes from ResFinder, we observed a drastic
decline in the number of drug classes a given isolate may present resistance to. The preva-
lence of tetracycline in human clinical and chicken production isolates is expected due to
its widespread usage in both agriculture and human clinical medicine and is of concern
because of its high importance to human medicine.
Of particular interest is our nding of widespread potential resistance to uoroqui-
nolone antibiotics within the chicken production isolates since their usage was banned
in poultry in 2005 (61), especially given that our data were obtained from samples
between 2018 and 2020. Fluoroquinolone is a critically important antimicrobial drug
class to human medicine due to its use in treatment of Campylobacter spp., invasive
Salmonella spp., and multidrug-resistant Shigella spp. infections (46). Further investiga-
tion of the causes for uoroquinolone resistance persistence in poultry is needed to
ensure its efcacy in treating human infections. Taken together, we identied many
antimicrobial drug classes which were highly present, but expected, in our study popu-
lation. However, the presence of AMR genes conferring resistance to antibiotics
banned for use in poultry, or food animal production in general, raises concern for their
future effectiveness in treating human infections and diseases.
We found 11 virulence functions, conferred by 251 identied VFs (Fig. 5B; see also Table
S14) to be present in both chicken and human clinical isolates. Based on hierarchical clus-
tering of T3SS genes, the E. coli isolates were classied into two distinct clusters, A and B.
The T3SS is associated with pathogens which can adhere to the epithelial surface of the
host, many of which can cause diarrheal disease; the second leading cause of death in chil-
dren globally (62, 63). Cluster A contained a higher representation of chicken production
isolates, while cluster B comprised mainly of human clinical isolates. This result suggests
that E. coli isolates from chicken production and human clinical isolates differ in their car-
riage of T3SS genes. The T3SS is essential for virulence and colonization of the human gut
and is the genetic basis for enteropathogenic E. coli classication (63). In addition, the T3SS
in poultry contributes to virulence in avian-pathogenic E. coli (64).Thisresultsuggeststhat
the T3SS genes can be used as targets for a rapid on-farm and clinical diagnostic detection
of virulent E. coli strains in chicken production and hospital settings. Identifying a lower
proportion of chicken production isolates harboring the T3SS is a positive nding since the
T3SS increases virulence in both humans and chickens.
Classication into host origin was successful with little error. Two of the most important
factors for this separation were the parC gene (S80I) and gyrA (S83L [n= 30], S84L [n=2],
and D87Y [n= 275]), which are both linked to uoroquinolone resistance. It has been docu-
mented that these resistance mutants are commonly found in environmental E. coli even
in the absence of uoroquinolone selective pressures (52). We speculate that these genes,
and their respective mutations, which relate to DNA replication and transcription, may con-
fer a tness advantage. The secreted enterotoxin TieB, encoded by senB,wasthesecond
most important factor in separating human and chicken isolates. senB was only identied
in human clinical isolates, and it was expected since senB is typically associated with
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 13
enteroinvasive and uropathogenic E. coli infection in humans (65). Similarly, TEM-1,a
broad-spectrum beta-lactamase conferring resistance to penicillins and rst-generation
cephalosporins, was also a high contributor to the separation of hosts and was only pres-
ent in human-linked isolates. While TEM-1 has been identied in poultry isolates (66), ceph-
alosporinusageismorecommoninbeefproduction(67).
This study reemphasizes the utility of phylogroup classication as a convenient way
to identify pathogenic strains. The most recent version of ClermonTyper (39) relies on
ve genes (arpA,chuA,yjaA, TspE4.C2, and ybgD) for phylogroup classication, only
one of the quadruplex genes, chuA (outer membrane protein responsible for heme
uptake), was present in our presence/absence table of identied AMR and VFs. Even
though yjaA (gene that differentiates phylogroups B2 and D isolates) was absent from
the AMR gene and VF table, we were able to reclassify B2 and D isolates with 100% ac-
curacy. Similarly, TspE4.C2 gene differentiates phylogroup A from phylogroup B1, and
its absence from our AMR gene and VF table resulted in an 11% classication error for
A isolates and a 3% error for B1 isolates. The absence of four of the ve ClermonTyping
genes might have contributed to the 48% error in classifying phylogroup C. The ability
to separate phylogroups F and G from B2 was recently accomplished with the addition
of the genes cfaB (CFA/I mbrial subunit B), specic to phylogroup G strains, and ybgD
(mbrial-like adhesin protein), specic to phylogroup F strains (39, 68). The absence of
these genes in our table only resulted in one phylogroup F (4% error rate) and one
phylogroup G misclassied isolates (2% error rate). Our results show that the phylotype
classication based on AMR and VFs could be achieved even when the majority of the
genes used for Clermont E. coli phylotyping are missing.
The use of whole-genome sequencing (WGS) is an effective tool to predict the AMR
and virulence potential of a bacterium; however, WGS still has its limitations. The depth
of coverage is crucial for genome assembly and gene annotation, both of which can
be signicantly inuenced by a coverage of less than 30. Our study excluded many
isolates from both chicken production and the human clinical settings due to low
sequencing coverage. Due to the nature of the database chosen for isolate selection,
i.e., a surveillance system for foodborne and other enteric bacteria, we expected an
over representation of isolates harboring AMR genes and VFs, and an under-represen-
tation of susceptible commensal and environmental E. coli isolates in our data sets.
In addition, the choice of AMR and virulence databases is important as not all databases
have the same entries and methods of detection. Here, we implemented Reads2Resistome
which allowed us to incorporate the results from multiple databases. Our nal AMR gene
and VFs were comprised of the highest percent hit based on similarity to the reference
gene (cutoff for gene coverage .80%) from each database. Another limitation of this study
is the lack of AMR or virulence phenotype data for each isolate. Therefore, the presence an
AMR gene or VFs in a genome is not a conrmation that the isolate is resistant to the anti-
biotic/drug class predicted or that the isolate can cause diseases.
MATERIALS AND METHODS
Study isolates, normalization, and assembly. A total of 1,711 Escherichia coli isolates, including
874 human clinical and 837 chicken production isolates, were selected from the USDAs GenomeTrackr
network via NCBIs Pathogen Detection Network. Agricultural poultry isolates were selected based on
the following criteria: species, E. coli and Shigella; location, USA; target creation, 01/01/2018 to 03/
30/2020; host, Poultry,”“poultry,”“Gallus gallus,and Gallus gallus domesticus.Human clinical isolates
were selected based on the following criteria: species, E. coli and Shigella; location, USA; target crea-
tion, 01/01/2018 to 03/30/2020; host, Homo sapiens,”“Homosapiens,”“homo sapiens,”“human,
human (infant).Only isolates with a straindenoted as E. coliwere retained for downstream analysis;
isolates identied as Shigellawere not retained. If available, Sequence Read Archive (SRA) reads, in
FASTQ format, were downloaded based on the BioSample ID for both agricultural and human isolates.
In the event the SRA reads were not available, FASTA assemblies were downloaded from the associated
BioProject submission. To ensure accurate and unbiased antimicrobial resistance identication, SRA
reads were further ltered based on coverage $20. To guarantee inclusion of quality genome assem-
blies, FASTA assemblies were ltered based on having an N
50
greater than 10,000 bp. Three isolates in
total were identied to be identical and were removed from downstream statistical analysis.
To compare genetic structure and contents, specically AMR genes, virulence factors, and plasmid
replicons across our study isolates, we assembled the SRA FASTQ reads into genomes, annotated these
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 14
genomes, and annotated the downloaded FASTA assemblies. To reduce bias in gene identication and
assembly quality across isolates, SRA reads were normalized to 30coverage using the insilico_read_-
normalization.pl script from Trinity RNA-Seq v2.11.0 (Trinitys GitHub). SRA reads were assembled using
Reads2Resistome v0.0.2 (options: slidingwindow = 4:20, assembly = nonhybrid, threads 60). Resulting
SRA assemblies were further ltered based on the N
50
(.10,000 bp) and nal coverage ($20) like the
FASTA assemblies described earlier. The nal study data set consisted of R2R assembled isolate genomes
and downloaded FASTA assemblies totaling 791 E. coli isolates (452 chicken production and 339 human
clinical; Fig. 2A).
Phylogenetic and sequence type analysis. To evaluate the relatedness of the isolates with respect
to their assembled genomes, we constructed a phylogenetic tree utilizing SNPs. These SNPs were identi-
ed by comparing the E. coli K12-MG1655 (accession U00096) reference genome to each of the 791 E.
coli isolates using Snippy v4.6.0 (69). The resulting whole-genome SNP alignment was then used to con-
struct a maximum-likelihood phylogenetic tree using the GTR1GAMMA model of evolution and utilizing
RAxML (RAxMLHPC-PTHREADS) v8.2.12 (70, 71). In addition, we performed 100 bootstraps to assess
nodal support. Only 100 bootstraps were performed due to time and computational limitations; it took
75.49 days utilizing 64 compute cores and 504 Gb memory to construct the phylogenetic tree and per-
form the 100 bootstraps. To determine which of the seven main phylogroups the isolates belong to, we
used ClermonTyping v1.4.0 (39), utilizing the default settings, to determine the phylogroup designation
(A, B1, B2, C, D, E, and F) for each isolate. Tree visualization, along with phylogroup and metadata over-
lay, was done in R v4.0.4 using ggtree v2.4.1 (72), factoextra v1.0.7 (73), and ggnewscale v0.4.5 (74). The
sequence type of each isolate was determined using ECTyper v1.0.0 (75) with default settings.
Antimicrobial resistance gene and virulence-factor characterization using Reads2Resistome.
Reads2Resistome (R2R; see the supplemental Materials and Methods) was used to characterize the antimicro-
bial resistance genes, virulence factors, and prophage regions. Reads2Resistome utilizes ABRICATE v0.5 (76)
to screen the assembled contigs against the following databases via BLAST and reports a percentage identity
to the reference sequence. Databases used by ABRICATE, and provided by R2R, were as follows: the ARG-
ANNOT antibiotic resistance gene database (77), the Comprehensive Antibiotic Resistance Database (CARD)
(45), the MEGARes Antimicrobial Database for High-Throughput Sequencing (78), NCBI AMRFinderPlus (79),
PlasmidFinder (80), and the VirulenceFinder database (81). Downloaded FASTA assembliesresistome charac-
terization was performed by the modied Reads2Resistome (R2R-0.0.1-Fasta-QC-Ann-Only.nf). In addition, re-
sistance gene identication was performed using the Resistance Gene Identier v5.1.1 (45) (RGI), and
acquired resistance genes were identied using ResFinder (82). AMR genes and virulence factors with the
highest percent identity across all databases were selected for further analysis. The heatmap of AMR genes
and virulence factors, using the table (described below; see also Table S1) was generated using pheatmap
v1.0.12 in R v4.0.4 with the average(unweighted pair group method with arithmetic mean [UPGMA]) clus-
tering method (see Fig. S1) (83). (See the supplemental Materials and Methods for further description of R2R).
Mobile genetic elements were determined for selected isolates using VRProle2 (updated most recently on 2
October 2021) (49).
Hierarchical clustering and statistical analysis of resistome and virulence factors. To evaluate
the relationships between poultry and human isolates with respect to their AMR and virulence factors,
we employed hierarchical clustering and ordination analysis. AMR and virulence database hits from R2R
and RGI were ltered utilizing $95% and $80% sequence identities to the reference database query,
respectively. ResFinder results were ltered utilizing $85% identity to the reference database query. A
table was then generated based on the presence/absence of identied VFs and AMR genes from all iso-
lates (see Table S1). A distance matrix was generated using the tanimoto metric via the Distance() (84)
function from the IntClust v0.1.0 (85) package. hclust() from the stats v3.6.2 package (86) was then uti-
lized to perform the hierarchical clustering under the average(UPGMA) method using the determined
optimal number of clusters. We identied the optimal number of clusters which separate the 791 iso-
lates using the silhouette method implemented by the fviz_nbclust() function from the factoextra v1.0.7
package (73). Drug classes associated with each identied AMR gene were determined using the output
from RGI, in combination with the CARD database (45), and virulence factor-associated functions were
determined using the comparative tables from the Virulence Factor Database (81). Genes conferring re-
sistance to drug classes were enumerated for each isolate, and a proportion was calculated using the
total number of genes in the study population conferring resistance to a given drug class. Drug classes
signicantly differing between hosts were determined by using a Wilcoxon rank sum test with the wil-
cox.test() function from the stats v3.6.2 package (86). Pvalue adjustment for multiple comparisons was
conducted using the p.adjust() function, according to the Benjamini and Hochberg (BH) method, from
the stats v3.6.2 package (86). PCA was conducted using the prcomp() function from the stats v3.6.2 pack-
age (86), and plots were generated using ggplot2 (87). All analyses were done in R v4.0.4 utilizing
RStudio v1.2.1106 (88).
Random forest classication. To determine the most inuential AMR and virulence factors driving
the separation of the isolates into (i) clusters identied via hierarchical clustering, (ii) isolate host origin,
and (iii) identied phylogroups, we employed random forest classication (89). The described presence/
absence table of AMR genes and virulence factors was used as input (see Table S1). TuneRF() (90) was
used to determine the optimal mtry parameter value, with respect to out-of-bag error estimate, for
RandomForest(). RandomForest() v4.6-14 (90) was used for classication with nttree = 200 based on the
ltered AMR and virulence-factor presence/absence table. All plots were generated using ggplot2 v3.3.3
(87). All analyses were done in R v4.0.4 utilizing RStudio v1.2.1106 (88).
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 15
Data availability. All raw isolate data, both FASTQ and FASTA les, were obtained from and are
available through the NCBI using the NCBI accession numbers in Table S1 in the supplemental material.
All other data are available upon request.
SUPPLEMENTAL MATERIAL
Supplemental material is available online only.
SUPPLEMENTAL FILE 1, PDF le, 0.9 MB.
SUPPLEMENTAL FILE 2, XLSX le, 0.9 MB.
SUPPLEMENTAL FILE 3, XLSX le, 0.01 MB.
ACKNOWLEDGMENTS
We are grateful to Jodie Plumblee Lawrence and Denice Cudnik for logistical and
technical assistance.
This study was supported by the U.S. Department of Agriculture (USDA) Agricultural
Research Service (project 6040-32000-012-000D). R.W. was supported by the National
Institutes of Health (grants 5T32GM132057-04 and 5T32GM132057-03). This research
was partially supported in part by an appointment to the Agricultural Research Service
(ARS) Research Participation Program administered by the Oak Ridge Institute for
Science and Education (ORISE) through an interagency agreement between the U.S.
Department of Energy (DOE) and the USDA (CRIS project 60-6040-6-009). ORISE is
managed by ORAU under DOE contract DE-SC0014664. All opinions expressed here are
the authorsand do not necessarily reect the policies and views of the USDA, the DOE,
or ORAU/ORISE. Any mention of products or trade names does not constitute
recommendation for use. We declare no competing commercial interests in relation to
the submitted work. The USDA is an equal opportunity provider and employer.
REFERENCES
1. Boyle DP, Zembower TR. 2015. Epidemiology and management of emerg-
ing drug-resistant Gram-negative bacteria: extended-spectrum
b
-lacta-
mases and beyond. Urol Clin North Am 42:493505. https://doi.org/10
.1016/j.ucl.2015.05.005.
2. Lynch M, Painter J, Woodruff R, Braden C, Centers for Disease Control and
Prevention. 2006. Surveillance for foodborne-disease outbreaksUnited
States, 19982002. MMWR Surveill Summ 55:142.
3. Dewey-Mattia D, Manikonda K, Hall AJ, Wise ME, Crowe SJ. 2018. Surveillance
for foodborne disease outbreaksUnited States, 20092015. MMWR Surveill
Summ 67:111. https://doi.org/10.15585/mmwr.ss6710a1.
4. Centers for Disease Control and Prevention. 2019. Surveillance for food-
borne disease outbreaks United States, 2014: annual report, p 116. Cen-
ters for Disease Control and Prevention, Atlanta, GA. http://www.cdc.gov/
foodsafety/fdoss/. Accessed 3 May 2022.
5. Russo TA, Johnson JR. 2003. Medical and economic impact of extraintesti-
nal infections due to Escherichia coli: focus on an increasingly important
endemic problem. Microbes Infect 5:449456. https://doi.org/10.1016/
S1286-4579(03)00049-2.
6. Nyachuba DG. 2010. Foodborne illness: is it on the rise? Nutr Rev 68:
257269. https://doi.org/10.1111/j.1753-4887.2010.00286.x.
7. Ramos S, Silva V, Dapkevicius MdLE, Caniça M, Tejedor-Junco MT, Igrejas
G, Poeta P. 2020. Escherichia coli as commensal and pathogenic bacteria
among food-producing animals: health implications of extended spec-
trum
b
-lactamase (ESBL) production. Animals (Basel) 10:2239. https://doi
.org/10.3390/ani10122239.
8. Croxen MA, Law RJ, Scholz R, Keeney KM, Wlodarska M, Finlay BB. 2013.
Recent advances in understanding enteric pathogenic Escherichia coli.
Clin Microbiol Rev 26:822880. https://doi.org/10.1128/CMR.00022-13.
9. Kaper JB, Nataro JP, Mobley HLT. 2004. Pathogenic Escherichia coli. Nat
Rev Microbiol 2:123140. https://doi.org/10.1038/nrmicro818.
10. Mainil J. 2013. Escherichia coli virulence factors. Vet Immunol Immunopa-
thol 152:212. https://doi.org/10.1016/j.vetimm.2012.09.032.
11. Khairy RM, Mohamed ES, Ghany HMA, Abdelrahim SS. 2019. Phylogenic clas-
sication and virulence genes proles of uropathogenic Escherichia coli and
diarrheagenic E. coli strains isolated from community acquired infections.
PLoS One 14:e0222441. https://doi.org/10.1371/journal.pone.0222441.
12. Yang SC, Lin CH, Aljuffali IA, Fang JY. 2017. Current pathogenic Escherichia
coli foodborne outbreak cases and therapy development. Arch Microbiol
199:811825. https://doi.org/10.1007/s00203-017-1393-y.
13. Panth Y. 2019. Colibacillosis in poultry: a review. J Agric Nat Res 2:
301311. https://doi.org/10.3126/janr.v2i1.26094.
14. Ewers C, Antão EM, Diehl I, Philipp HC, Wieler LH. 2009. Intestine and envi-
ronment of the chicken as reservoirs for extraintestinal pathogenic Esche-
richia coli strains with zoonotic potential. Appl Environ Microbiol 75:
184192. https://doi.org/10.1128/AEM.01324-08.
15. Dho-Moulin M, Morris Fairbrother J. 1999. Avian pathogenic Escherichia
coli (APEC). Vet Res 30(2-3):299366.
16. Guabiraba R, Schouler C. 2015. Avian colibacillosis: still many black holes.
FEMS Microbiol Lett 362:fnv118. https://doi.org/10.1093/femsle/fnv118.
17. Mellata M. 2013. Human and avian extraintestinal pathogenic Escherichia
coli: infections, zoonotic risks, and antibiotic resistance trends. Foodborne
Pathog Dis 10:916932. https://doi.org/10.1089/fpd.2013.1533.
18. Johnson TJ, Siek KE, Johnson SJ, Nolan LK. 2006. DNA sequence of a ColV
plasmid and prevalence of selected plasmid-encoded virulence genes
among avian Escherichia coli strains. J Bacteriol 188:745758. https://doi
.org/10.1128/JB.188.2.745-758.2006.
19. Mitchell NM, Johnson JR, Johnston B, Curtiss R, Mellata M. 2015. Zoonotic
potential of Escherichia coli isolates from retail chicken meat products
and eggs. Appl Environ Microbiol 81:11771187. https://doi.org/10.1128/
AEM.03524-14.
20. Johnson TJ, Wannemuehler Y, Johnson SJ, Stell AL, Doetkott C, Johnson
JR, Kim KS, Spanjaard L, Nolan LK. 2008. Comparison of extraintestinal
pathogenic Escherichia coli strains from human and avian sources reveals
a mixed subset representing potential zoonotic pathogens. Appl Environ
Microbiol 74:70437050. https://doi.org/10.1128/AEM.01395-08.
21. Johnson TJ, Kariyawasam S, Wannemuehler Y, Mangiamele P, Johnson SJ,
Doetkott C, Skyberg JA, Lynne AM, Johnson JR, Nolan LK. 2007. The genome
sequence of avian pathogenic Escherichia coli strain O1:K1:H7 shares strong
similarities with human extraintestinal pathogenic E. coli genomes. J Bacteriol
189:32283236. https://doi.org/10.1128/JB.01726-06.
22. Moulin-Schouleur M, Schouler C, Tailliez P, Kao M-R, Brée A, Germon P,
Oswald E, Mainil J, Blanco M, Blanco J. 2006. Common virulence factors
and genetic relationships between O18:K1:H7 Escherichia coli isolates of
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 16
human and avian origin. J Clin Microbiol 44:34843492. https://doi.org/
10.1128/JCM.00548-06.
23. EwersC,LiG,WilkingH,KiesslingS,AltK,AntáoE-M,LaturnusC,DiehlI,
Glodde S, Homeier T, Böhnke U, Steinrück H, Philipp H-C, Wieler LH. 2007.
Avian pathogenic, uropathogenic, and newborn meningitis-causing Esche-
richia coli: how closely related are they? Int J Med Microbiol 297:163176.
https://doi.org/10.1016/j.ijmm.2007.01.003.
24. JakobsenL,GarneauP,KurbasicA,BruantG,SteggerM,HarelJ,JensenKS,
Brousseau R, Hammerum AM, Frimodt-Møller N. 2011. Microarray-based detec-
tion of extended virulence and antimicrobial resistance gene proles in phy-
logroup B2 Escherichia coli of human, meat and animal origin. J Med Microbiol
60:15021511. https://doi.org/10.1099/jmm.0.033993-0.
25.TivendaleKA,LogueCM,KariyawasamS,JordanD,HusseinA,LiG,
Wannemuehler Y, Nolan LK. 2010. Avian-pathogenic Escherichia coli strains
are similar to neonatal meningitis E. coli strainsandareabletocausemenin-
gitis in the rat model of human disease. Infect Immun 78:34123419. https://
doi.org/10.1128/IAI.00347-10.
26. Comery R, Thanabalasuriar A, Garneau P, Portt A, Boerlin P, Reid-Smith RJ,
Harel J, Manges AR, Gruenheid S. 2013. Identication of potentially diar-
rheagenic atypical enteropathogenic Escherichia coli strains present in Ca-
nadian food animals at slaughter and in retail meats. Appl Environ Micro-
biol 79:38923896. https://doi.org/10.1128/AEM.00182-13.
27. Jakobsen L, Garneau P, Bruant G, Harel J, Olsen SS, Porsbo LJ, Hammerum
AM, Frimodt-Møller N. 2012. Is Escherichia coli urinary tract infection a zoono-
sis? Proof of direct link with production animals and meat. Eur J Clin Micro-
biol Infect Dis 31:11211129. https://doi.org/10.1007/s10096-011-1417-5.
28. Li G, Cai W, Hussein A, Wannemuehler YM, Logue CM, Nolan LK. 2012.
Proteome response of an extraintestinal pathogenic Escherichia coli strain
with zoonotic potential to human and chicken sera. J Proteomics 75:
48534862. https://doi.org/10.1016/j.jprot.2012.05.044.
29. Skyberg JA, Johnson TJ, Johnson JR, Clabots C, Logue CM, Nolan LK. 2006.
Acquisition of avian pathogenic Escherichia coli plasmids by a commensal
E. coli isolate enhances its abilities to kill chicken embryos, grow in human
urine, and colonize the murine kidney. Infect Immun 74:62876292.
https://doi.org/10.1128/IAI.00363-06.
30. Caniça M, Manageiro V, Abriouel H, Moran-Gilad J, Franz CMAP. 2019. An-
tibiotic resistance in foodborne bacteria. Trends Food Sci Technol 84:
4144. https://doi.org/10.1016/j.tifs.2018.08.001.
31. Ferri M, Ranucci E, Romagnoli P, Giaccone V. 2017. Antimicrobial resist-
ance: a global emerging threat to public health systems. Crit Rev Food Sci
Nutr 57:28572876. https://doi.org/10.1080/10408398.2015.1077192.
32. Fenske GJ, Scaria J. 2021. Analysis of 56,348 genomes identies the rela-
tionship between antibiotic and metal resistance and the spread of multi-
drug-resistant non-typhoidal Salmonella. Microorganisms 9:1468. https://
doi.org/10.3390/microorganisms9071468.
33. Oladeinde A, Abdo Z, Zwirzitz B, Woyda R, Lakin SM, Press MO, Cox NA,
Thomas JC, Looft T, Rothrock MJ, Zock G, Plumblee Lawrence J, Cudnik D,
Ritz C, Aggrey SE, Liachko I, Grove JR, Wiersma C. 2022. Litter commensal
bacteria can limit the horizontal gene transfer of antimicrobial resistance
to Salmonella in chickens. Appl Environ Microbiol 88:e02517-22. https://
doi.org/10.1128/aem.02517-21.
34. Biswas K, Sharma P, Joshi SR. 2019. Co-occurrence of antimicrobial resist-
ance and virulence determinants in enterococci isolated from tradition-
ally fermented sh products. J Glob Antimicrob Resist 17:7983. https://
doi.org/10.1016/j.jgar.2018.11.012.
35. Pan Y, Zeng J, Li L, Yang J, Tang Z, Xiong W, Li Y, Chen S, Zeng Z. 2020.
Coexistence of antibiotic resistance genes and virulence factors deci-
phered by large-scale complete genome analysis. mSystems 5:e00821-19.
https://doi.org/10.1128/mSystems.00821-19.
36. Gurevich A, Saveliev V, Vyahhi N, Tesler G. 2013. QUAST: quality assess-
ment tool for genome assemblies. Bioinformatics 29:10721075. https://
doi.org/10.1093/bioinformatics/btt086.
37. Woyda R, Oladeinde A, Abdo Z. 2020. Reads2Resistome: an adaptable and
high-throughput whole-genome sequencing pipeline for bacterial resistome
characterization. bioRxiv. https://www.biorxiv.org/content/10.1101/2020.05
.18.102715v1.
38. Kurtzer GM, Sochat V, Bauer MW. 2017. Singularity: scientic containers
for mobility of compute. PLoS One 12:e0177459. https://doi.org/10.1371/
journal.pone.0177459.
39. Clermont O, Dixit OVA, Vangchhia B, Condamine B, Dion S, Bridier-
Nahmias A, Denamur E, Gordon D. 2019. Characterization and rapid iden-
tication of phylogroup G in Escherichia coli, a lineage with high virulence
and antibiotic resistance potential. Environ Microbiol 21:31073117.
https://doi.org/10.1111/1462-2920.14713.
40. Clermont O, Christenson JK, Denamur E, Gordon DM. 2013. The Clermont
Escherichia coli phylotyping method revisited: improvement of specicity
and detection of new phylogroups. Environ Microbiol Rep 5:5865. https://
doi.org/10.1111/1758-2229.12019.
41. Ingle DJ, Clermont O, Skurnik D, Denamur E, Walk ST, Gordon DM. 2011.
Biolm formation by and thermal niche and virulence characteristics of
Escherichia spp. Appl Environ Microbiol 77:26952700. https://doi.org/10
.1128/AEM.02401-10.
42. Smati M, Clermont O, Le Gal F, Schichmanoff O, Jauréguy F, Eddi A, Denamur
E, Picard B, Coliville Group. 2013. Real-time PCR for quantitative analysis of
human commensal Escherichia coli populations reveals a high frequency of
subdominant phylogroups. Appl Environ Microbiol 79:50055012. https://
doi.org/10.1128/AEM.01423-13.
43. Ahumada-Santos YP, Báez-Flores ME, Díaz-Camacho SP, Uribe-Beltrán
MdJ, Eslava-Campos CA, Parra-Unda JR, Delgado-Vargas F. 2020. Associa-
tion of phylogenetic distribution and presence of integrons with multi-
drug resistance in Escherichia coli clinical isolates from children with diar-
rhoea. J Infect Public Health 13:767772. https://doi.org/10.1016/j.jiph
.2019.11.019.
44. Zankari E, Hasman H, Cosentino S, Vestergaard M, Rasmussen S, Lund O,
Aarestrup FM, Larsen MV. 2012. Identication of acquired antimicrobial
resistance genes. J Antimicrob Chemother 67:26402644. https://doi.org/
10.1093/jac/dks261.
45. Alcock BP, Raphenya AR, Lau TTY, Tsang KK, Bouchard M, Edalatmand A,
Huynh W, Nguyen A-LV, Cheng AA, Liu S, Min SY, Miroshnichenko A, Tran
H-K, Werfalli RE, Nasir JA, Oloni M, Speicher DJ, Florescu A, Singh B, Faltyn
M, Hernandez-Koutoucheva A, Sharma AN, Bordeleau E, Pawlowski AC,
Zubyk HL, Dooley D, Grifths E, Maguire F, Winsor GL, Beiko RG, Brinkman
FSL, Hsiao WWL, Domselaar GV, McArthur AG. 2020. CARD 2020: antibi-
otic resistome surveillance with the comprehensive antibiotic resistance
database. Nucleic Acids Res 48:D517D525. https://doi.org/10.1093/nar/
gkz935.
46. WHO. 2019. WHO list of critically important antimicrobials (CIA). World
Health Organization, Geneva, Switzerland.
47. Bagel S, Hüllen V, Wiedemann B, Heisig P. 1999. Impact of gyrA and parC
mutations on quinolone resistance, doubling time, and supercoiling degree
of Escherichia coli. Antimicrob Agents Chemother 43:868875. https://doi
.org/10.1128/AAC.43.4.868.
48. Girgis HS, Hottes AK, Tavazoie S. 2009. Genetic architecture of intrinsic an-
tibiotic susceptibility. PLoS One 4:e5629. https://doi.org/10.1371/journal
.pone.0005629.
49. Wang M, Goh Y-X, Tai C, Wang H, Deng Z, Ou H-Y. 2022. VRprole2: detec-
tion of antibiotic resistance-associated mobilome in bacterial pathogens.
Nucleic Acids Res 50:W768W773. https://doi.org/10.1093/nar/gkac321.
50. Tseytin I, Dagan A, Oren S, Sal-Man N. 2018. The role of EscD in supporting
EscC polymerization in the type III secretion system of enteropathogenic
Escherichia coli. Biochim Biophys Acta Biomembr 1860:384395. https://doi
.org/10.1016/j.bbamem.2017.10.001.
51. Slater SL, Sågfors AM, Pollard DJ, Ruano-Gallego D, Frankel G. 2018. The
type III secretion system of pathogenic Escherichia coli. Curr Top Microbiol
Immunol 416:5172. https://doi.org/10.1007/82_2018_116.
52. Tavío MM, Vila J, Ruiz J, Ruiz J, Martín-Sánchez AM, Jiménez de Anta MT. 1999.
Mechanisms involved in the development of resistance to uoroquinolones
in Escherichia coli isolates. J Antimicrob Chemother 44:735742. https://doi
.org/10.1093/jac/44.6.735.
53. Johnning A, Kristiansson E, Fick J, Weijdegård B, Larsson DGJ. 2015. Resist-
ance mutations in gyrA and parC are common in Escherichia communities
of both uoroquinolone-polluted and uncontaminated aquatic environ-
ments. Front Microbiol 6:1355. https://doi.org/10.3389/fmicb.2015.01355.
54. Sáenz Y, Zarazaga M, Briñas L, Ruiz-Larrea F, Torres C. 2003. Mutations in gyrA
and parC genes in nalidixic acid-resistant Escherichia coli strains from food
products, humans and animals. J Antimicrob Chemother 51:10011005.
https://doi.org/10.1093/jac/dkg168.
55. Gao Q, Wang X, Xu H, Xu Y, Ling J, Zhang D, Gao S, Liu X. 2012. Roles of
iron acquisition systems in virulence of extraintestinal pathogenic Esche-
richia coli: salmochelin and aerobactin contribute more to virulence than
heme in a chicken infection model. BMC Microbiol 12:143. https://doi.org/
10.1186/1471-2180-12-143.
56. Sorsa LJ, Dufke S, Heesemann J, Schubert S. 2003. Characterization of an
iroBCDEN gene cluster on a transmissible plasmid of uropathogenic Esch-
erichia coli: evidence for horizontal transfer of a chromosomal virulence
factor. Infect Immun 71:32853293. https://doi.org/10.1128/IAI.71.6.3285
-3293.2003.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 17
57. Katouli M. 2010. Population structure of gut Escherichia coli and its role in
development of extra-intestinal infections. Iran J Microbiol 2:5972.
58. CortésP,BlancV,MoraA,DahbiG,BlancoJE,BlancoM,LópezC,AndreuA,
Navarro F, Alonso MP, Bou G, Blanco J, Llagostera M. 2010. Isolation and char-
acterization of potentially pathogenic antimicrobial-resistant Escherichia coli
strains from chicken and pig farms in Spain. Appl Environ Microbiol 76:
27992805. https://doi.org/10.1128/AEM.02421-09.
59. Van Hoek AHAM, Mevius D, Guerra B, Mullany P, Roberts AP, Aarts HJM.
2011. Acquired antibiotic resistance genes: an overview. Front Microbiol
2:203. https://doi.org/10.3389/fmicb.2011.00203.
60. García MR, Cabo ML. 2018. Optimization of Escherichia coli inactivation by
benzalkonium chloride reveals the importance of quantifying the inocu-
lum effect on chemical disinfection. Front Microbiol 9:1259. https://doi
.org/10.3389/fmicb.2018.01259.
61. Price LB, Lackey LG, Vailes R, Silbergeld E. 2007. The persistence of uoro-
quinolone-resistant campylobacter in poultry production. Environ Health
Perspect 115:10351039. https://doi.org/10.1289/ehp.10050.
62.BlackRE,CousensS,JohnsonHL,LawnJE,RudanI,BassaniDG,JhaP,
Campbell H, Walker CF, Cibulskis R, Eisele T, Liu L, Mathers C, Child Health Ep-
idemiology Reference Group of WHO and UNICEF. 2010. Global, regional,
and national causes of child mortality in 2008: a systematic analysis. Lancet
375:19691987. https://doi.org/10.1016/S0140-6736(10)60549-1.
63. Deborah Chen H, Frankel G. 2005. Enteropathogenic Escherichia coli:
unraveling pathogenesis. FEMS Microbiol Rev 29:8398. https://doi.org/
10.1016/j.femsre.2004.07.002.
64. Wang S, Liu X, Xu X, Yang D, Wang D, Han X, Shi Y, Tian M, Ding C, Peng
D, Yu S. 2016. Escherichia coli type III secretion system 2 ATPase EivC is
involved in the motility and virulence of avian pathogenic Escherichia coli.
Front Microbiol 7:1387. https://doi.org/10.3389/fmicb.2016.01387.
65. Mao B-H, Chang Y-F, Scaria J, Chang C-C, Chou L-W, Tien N, Wu J-J, Tseng
C-C, Wang M-C, Chang C-C, Hsu Y-M, Teng C-H. 2012. Identication of
Escherichia coli genes associated with urinary tract infections. J Clin Micro-
biol 50:449456. https://doi.org/10.1128/JCM.00640-11.
66. Seo KW, Lee YJ. 2018. Prevalence and characterization of
b
-lactamases
genes and class 1 integrons in multidrug-resistant Escherichia coli isolates
from chicken meat in Korea. Microb Drug Resist 24:15991606. https://
doi.org/10.1089/mdr.2018.0019.
67. FDA. 2018. Summary report on antimicrobials sold or distributed for use
in food-producing animals. Cent Vet Med 2018:114. https://www.fda
.gov/media/144427/download.
68. Lu S, Jin D, Wu S, Yang J, Lan R, Bai X, Liu S, Meng Q, Yuan X, Zhou J, Pu J,
Chen Q, Dai H, Hu Y, Xiong Y, Ye C, Xu J. 2016. Insights into the evolution
of pathogenicity of Escherichia coli from genomic analysis of intestinal E.
coli of Marmota himalayana in QinghaiTibet plateau of China. Emerg
Microbes Infect 5:e122. https://doi.org/10.1038/emi.2016.122.
69. Seemann T. 2015. Snippy, rapid haploid variant calling and core genome
alignment. GitHub Repository https://github.com/tseemann/snippy.
70. Lutteropp S, Kozlov AM, Stamatakis A. 2019. A fast and memory-efcient
implementation of the transfer bootstrap. bioRxiv. https://www.biorxiv
.org/content/10.1101/734848v2.
71. Stamatakis A. 2006. RAxML-VI-HPC: maximum likelihood-based phyloge-
netic analyses with thousands of taxa and mixed models. Bioinformatics
22:26882690. https://doi.org/10.1093/bioinformatics/btl446.
72. Yu G. 2020. Using ggtree to visualize data on tree-like structures. Curr Pro-
toc Bioinform 69:e96. https://doi.org/10.1002/cpbi.96.
73. Kassambara A, Mundt F. 2020. Factoextra: extract and visualize the results of
multivariate data analyses. R Package Version 1.0.7. https://CRAN.R-project
.org/package=factoextra.
74. Campitelli E. 2022. ggnewscale: multiple ll and colour scales in ggplot2.
https://doi.org/10.5281/zenodo.2543762.
75. Bessonov K, Laing C, Robertson J, Yong I, Ziebell K, Gannon VPJ, Nichani
A, Arya G, Nash JHE, Christianson S. 2021. ECTyper: in silico Escherichia coli
serotype and species prediction from raw and assembled whole-genome
sequence data. Microb Genom 7:000728. https://doi.org/10.1099/mgen.0
.000728.
76. Seemann T. 2020. Abricate. Github https://github.com/tseemann/abricate.
77. Gupta SK, Padmanabhan BR, Diene SM, Lopez-Rojas R, Kempf M, Landraud L,
Rolain J-M. 2014. ARG-annot, a new bioinformatic tool to discover antibiotic
resistance genes in bacterial genomes. Antimicrob Agents Chemother 58:
212220. https://doi.org/10.1128/AAC.01310-13.
78. Lakin SM, Dean C, Noyes NR, Dettenwanger A, Ross AS, Doster E, Rovira P,
Abdo Z, Jones KL, Ruiz J, Belk KE, Morley PS, Boucher C. 2017. MEGARes:
an antimicrobial resistance database for high throughput sequencing.
Nucleic Acids Res 45:D574D580. https://doi.org/10.1093/nar/gkw1009.
79. Feldgarden M, Brover V, Haft DH, Prasad AJ, Slotta DJ, et al. 2019. Validating
the AMRFinder tool and resistance gene database by using antimicrobial re-
sistance genotype-phenotype correlations in a collection of isolates. Antimi-
crob Agents Chemother 63:e00483-19. https://doi.org/10.1128/AAC.00483-19.
80. Carattoli A, Zankari E, García-Fernández A, Voldby Larsen M, Lund O, Villa L,
Møller Aarestrup F, Hasman H. 2014. In silico detection and typing of plasmids
using plasmidnder and plasmid multilocus sequence typing. Antimicrob
Agents Chemother 58:38953903. https://doi.org/10.1128/AAC.02412-14.
81. Liu B, Zheng D, Jin Q, Chen L, Yang J. 2019. VFDB 2019: a comparative
pathogenomic platform with an interactive web interface. Nucleic Acids
Res 47:D687D692. https://doi.org/10.1093/nar/gky1080.
82. Bortolaia V, Kaas RS, Ruppe E, Roberts MC, Schwarz S, Cattoir V, Philippon A,
Allesoe RL, Rebelo AR, Florensa AF, Fagelhauer L, Chakraborty T, Neumann B,
Werner G, Bender JK, Stingl K, Nguyen M, Coppens J, Xavier BB, Malhotra-
Kumar S, Westh H, Pinholt M, Anjum MF, Duggett NA, Kempf I, Nykäsenoja S,
Olkkola S, Wieczorek K, Amaro A, Clemente L, Mossong J, Losch S, Ragimbeau
C, Lund O, Aarestrup FM. 2020. ResFinder 4.0 for predictions of phenotypes
from genotypes. J Antimicrob Chemother 75:34913500. https://doi.org/10
.1093/jac/dkaa345.
83. Sokal RR. 1958. A statistical method for evaluating systematic relationships.
Univ Kansas Sci Bull 38:14091438. https://doi.org/10.24517/00055549.
84. Miller DL, Rexstad E, Thomas L, Marshall L, Laake JL. 2019. Distance sam-
pling in R. J Statist Softw 89:128. https://doi.org/10.18637/jss.v089.i01.
85. Moerbeke MV. 2018. IntClust: integration of multiple data sets with clus-
tering techniques. Version 0.1.0. https://CRAN.R-project.org/package=
IntClust.
86. R Core Team. 2015. R: a language and environment for statistical comput-
ing. R Foundation for Statistical Computing, Vienna, Austria. https://www
.R-project.org.
87. Wickham H. 2016. ggplot2: elegant graphics for data analysis. Springer-Ver-
lag, New York, NY. ISBN 9783-31924277-4. https://ggplot2.tidyverse.org.
88. RStudio, Inc. 2016. Integrated development environment for R. RStudio,
Inc, Boston, MA.
89. Breiman L. 2001. Random forests. Machine Learning 45:532. https://doi
.org/10.1023/A:1010933404324.
90. Liaw A, Wiener M. 2002. Classication and regression by randomForest. R
Newsl 2:1822. https://CRAN.R-project.org/doc/Rnews/.
Chicken and Human E. coli Differ in AMR and Virulence Applied and Environmental Microbiology
February 2023 Volume 89 Issue 2 10.1128/aem.01167-22 18
... Historically, the most prevalent AMR genes detected in E. coli from retail meat from the NARMS surveillance program include genes associated with erythromycin (mph(A)), tetracycline (tet(A), tet(B) and tet(C)), sulfonamide (sul1 and sul2) and plasmid mediated quinolone (qnr; gyr or par mutations) resistance [16][17][18]. Recent analysis of the NARMS Genome Trackr database showed that common phylogroups detected in E. coli recovered from retail meat were A, B1, B2, C, Clade I, D, E, F, and G [18]. ...
... Historically, the most prevalent AMR genes detected in E. coli from retail meat from the NARMS surveillance program include genes associated with erythromycin (mph(A)), tetracycline (tet(A), tet(B) and tet(C)), sulfonamide (sul1 and sul2) and plasmid mediated quinolone (qnr; gyr or par mutations) resistance [16][17][18]. Recent analysis of the NARMS Genome Trackr database showed that common phylogroups detected in E. coli recovered from retail meat were A, B1, B2, C, Clade I, D, E, F, and G [18]. Previous NARMS studies have reported E. coli prevalence of 47.5% in all retail meat products, with higher prevalence (90.7%) reported in turkey products [17,19]. ...
... We found the presence of AMR genes against different antimicrobial drug classes (aminoglycosides, tetracyclines, beta-lactamases, folate pathway antagonists, quinolones, and macrolides) in E. coli isolates from chicken, turkey and pork sources, but the prevalence differed by the various retail meat types. Our study results are consistent with historical genotypic prevalence reported in previous NARMS surveillance program [16][17][18]. ...
Article
Full-text available
Background Escherichia coli is commonly used as an indicator for antimicrobial resistance (AMR) in food, animal, environment, and human surveillance systems. Our study aimed to characterize AMR in E . coli isolated from retail meat purchased from grocery stores in North Carolina, USA as part of the National Antimicrobial Resistance Monitoring System (NARMS). Materials and methods Retail chicken (breast, n = 96; giblets, n = 24), turkey (n = 96), and pork (n = 96) products were purchased monthly from different counties in North Carolina during 2022. Label claims on packages regarding antibiotic use were recorded at collection. E . coli was isolated from meat samples using culture-based methods and isolates were characterized for antimicrobial resistance using whole genome sequencing. Multi-locus sequence typing, phylogroups, and a single nucleotide polymorphism (SNP)-based maximum-likelihood phylogenic tree was generated. Data were analyzed statistically to determine differences between antibiotic use claims and meat type. Results Of 312 retail meat samples, 138 (44.2%) were positive for E . coli , with turkey (78/138; 56.5%) demonstrating the highest prevalence. Prevalence was lower in chicken (41/138; 29.7%) and pork (19/138;13.8%). Quality sequence data was available from 84.8% (117/138) of the E . coli isolates, which included 72 (61.5%) from turkey, 27 (23.1%) from chicken breast, and 18 (15.4%) from pork. Genes associated with AMR were detected in 77.8% (91/117) of the isolates and 35.9% (42/117) were defined as multidrug resistant (MDR: being resistant to ≥3 distinct classes of antimicrobials). Commonly observed AMR genes included tetB (35%), tetA (24.8%), aph(3’’)-lb (24.8%), and bla TEM-1 (20.5%), the majority of which originated from turkey isolates. Antibiotics use claims had no statistical effect on MDR E . coli isolates from the different meat types ( X ² = 2.21, p = 0.33). MDR was observed in isolates from meat products with labels indicating “no claims” (n = 29; 69%), “no antibiotics ever” (n = 9; 21.4%), and “organic” (n = 4; 9.5%). Thirty-four different replicon types were observed. AMR genes were carried on plasmids in 17 E . coli isolates, of which 15 (88.2%) were from turkey and two (11.8%) from chicken. Known sequence types (STs) were described for 81 E . coli isolates, with ST117 (8.5%), ST297 (5.1%), and ST58 (3.4%) being the most prevalent across retail meat types. The most prevalent phylogroups were B1 (29.1%) and A (28.2%). Five clonal patterns were detected among isolates. Conclusions E . coli prevalence and the presence of AMR and MDR were highest in turkey retail meat. The lack of an association between MDR E . coli in retail meat and antibiotic use claim, including those with no indication of antimicrobial use, suggests that additional research is required to understand the origin of resistance. The presence of ST117, an emerging human pathogen, warrants further surveillance. The isolates were distinctly diverse suggesting an instability in population dynamics.
... Moreover, it is necessary to realize that humans are part of the ecosystem and are constantly in fluenced by other components of the ecosystem, especially as anthropogenic influence on the environment is unquestionable. It cannot be denied that people use chemicals, including antibiotics, in agriculture, fish farming, and animal husbandry to obtain richer harvests or better foodproducing animals to compete on the market (12,13,14). Farms and animal husbandry are important sources of antimicrobial-resistant pathogens and an important component that ensures the continuous AMR cycle in the environment, due to animal care and handling processes including treatment, hygiene, and slaughter (14 -17). ...
... The limited number of methods used to combat AMR, combined with the preference of LMIC patients (and others) towards self-medication, are factors that make it difficult to align with the 2015 World Health Assembly Global Action Plan on AMR goals, particularly goals 4 and 5 (2,3,22,23,37). The overuse of antibiotics in LMICs, which has increased by 65% over the last decade, resulting in the emergence of multi-drug resistant (MDR) and extensively drug-resistant (XDR) superbugs, calls for innovative measures to combat AMR at the level of every component of the ecosystem (5,7,14,19,24). The European Union also emphasizes the role of Gram-negative bacteria in the etiology of infections with antimicrobial-resistant bacteria, which supports the results of World Health Organization (WHO) reviews concerning AMR (25). New antimicrobials and combinations are being sought and developed to combat MDR Enterobacteriaceae (e.g. ...
Article
Full-text available
Introduction. Antimicrobial resistance has been declared a major public health problem. As a result of excesive antibiotic usage, it became an environmental issue. This problem is now more visible in Low-and Middle-Income Countries, where it increases the social burden. One of the newest methods to fight antimicrobial-resistant bacteria (ARB) is by using bacteriophages. Aim Identification and characterization of possible phage targets from waterland. Material and methods The strains were obtained from inpatients and identified using VITEK 2 Compact and culture. The resistance profiles were determined by disk diffusion method and interpreted according to EUCAST metodology. The presence of resistance mechanisms was checked by phenotypic testing. 31 bacterial strains were selected for research. Results K. pneumoniae, P. aeruginosa, Acinetobacter spp., S. aureus, E. coli and Enterococcus spp. were identified. The resistance profile of the isolates showed: 61,5% of K. pneumoniae isolates were PDR, and 23,1% were susceptible only to Carbapenems. E. coli strains were XDR, 71,4% of P. aeruginosa and 75% of Acinetobacter spp. were PDR bacteria. The susceptibility profile of S. aureus strains showed that 3/4 were resistant to Cephalosporins and Fluoroquinolones. Conclusions Combating the AMR phenomenon starts with knowing the pathogens present in the environment. This study is the cornerstone for further research that will ultimately lead to finding suitable phages for water treatment in Moldova, hoping they will reduce the economic and social burden and also, environmental contamination.
... In terms of public health, these pathogenic poultry bacteria may induce fever, stomach cramps, diarrhea, and nausea, whereas infections by Listeria spp. can also induce stiff neck, confusion, loss of balance, convulsions, hospitalization, and fatality [5][6][7]. Poultry products are among the most common vehicles for human transmission of Salmonella spp. and Campylobacter spp., with the latter accounting for the majority of bacterial gastroenteritis cases in Europe [8]. ...
Article
Full-text available
The aim of the study was to investigate in vitro the antibacterial activity of 8 commercial drinking water additives against major zoonotic poultry pathogens (Campylobacter spp., Escherichia coli, Salmonella Typhimurium, Staphylococcus aureus and Listeria spp.). We tested two essential oil-based phytogenics (Phyto CSC Liquide B, AEN 350 B Liquid), two acid-based eubiotics (Salgard® liquid, Intesti-Flora), and four blends of essential oils and organic acids (ProPhorceTM SA Exclusive, Herbal acid, Rigosol-N and Eubisan 3000). The antibacterial activity was determined by estimating the minimum inhibitory concentration (MIC) using a microdilution method. The MICs of the products against Campylobacter spp. ranged from 0.071% to 0.568% v/v, in which Herbal acid, a blend rich in lactic and phosphoric acids, also containing thyme and oregano oils, exhibited the highest efficacy (MIC: 0.071% v/v) against all the tested strains. The MICs of the tested products against Escherichia coli ranged between 0.071% and 1.894% v/v. Specifically, the MIC of Rigosol-N, a blend of high concentrations of lactic and acetic acid, was 0.142% v/v for both tested strains, whereas the MICs of Intesti-Flora, a mixture rich in lactic and propionic acid, ranged from 0.284% to 0.568% v/v. The MICs of the products against Salmonella Typhimurium were between 0.095% and 1.894% v/v. Specifically, the MIC of Eubisan 3000, a blend rich in oregano oil, was 0.284% v/v. The MICs against Staphylococcus aureus were between 0.142% and 9.090% v/v. The MICs of Phyto CSC Liquide B, which is rich in trans-cinnamaldehyde, were between 3.030% and 9.090% v/v, showing the highest MIC values of all tested products. Finally, the MIC values of the tested commercial products against Listeria spp. were 0.095% to 3.030% v/v. The MICs of ProPhorceTM SA Exclusive, a highly concentrated blend of formic acid and its salts, were 0.095–0.142% v/v against Listeria spp., while the MICs of AEN 350 B Liquid were between 0.284% and 1.894% exhibiting high Listeria spp. strain variability. In conclusion, all the selected commercial products exhibited more or less antibacterial activity against pathogenic bacteria and, thus, can be promising alternatives to antibiotics for the control of zoonotic poultry pathogens and the restriction of antimicrobial-resistant bacteria.
... Resistance [50], and Germany (3.7%) [51]. According to Woyda et al. [52], chicken production isolates were more likely than human clinical isolates to carry acquired AMR genes (65.7%), and they also carried an average of more acquired AMR genes per isolate. Furthermore, Juraschek et al. [51] verified that the existence of qnr alone can result in phenotypic (fluoro) QR without the need for PMQR or point mutations in the relevant chromosomal area. ...
Article
Full-text available
Objective The goal of this study was to look at quinolone-resistant (QR) Escherichia coli (E. coli) from retail beef and poultry meat in Egypt by looking at the QR mechanisms in the resistant strains. Materials and Methods In total, 120 samples of raw poultry meat (n = 60) and beef meat (n = 60) were purchased from Mansoura retail stores between January and March 2021, and evaluated microbiologically for E. coli. Then, an antimicrobial sensitivity test was applied to all isolates. The prevalence of QR E. coli with concern for the QR determinants, including quinolone resistance-determining regions (QRDRs) mutations, the plasmid-mediated quinolone resistance gene (PMQR), and the efflux pump activity were determined. Results The total prevalence of E. coli was 34.2% (41/120). Noticeably, the prevalence of E. coli in poultry meat (40%, 24/60) was higher than that of beef (28%, 17/60). All strains were assessed for their antimicrobial susceptibility using the disc diffusion technique; the highest rate of resistance (100%) was displayed to clindamycin and cefuroxime, followed by ampicillin (97.6%), doxycycline (92.7%), amoxicillin-clavulanate (92.7%), nalidixic acid (NA) (80.5%), sulfamethoxazole/trimethoprim (70.7%), chloramphenicol (63.4%), gentamicin, and azithromycin (58.5% each). Multiple antimicrobial resistance (strains resistant to three or more antimicrobial classes) was displayed by 97.6% of E. coli isolates. Regarding QR, 37 isolates could resist at least one of the examined quinolones. Regarding PMQR genes, qnrS was determined in 70% (7/10) of QR E. coli, while qnrA, qnrB, and qnrD were not identified. While the mutations determined regions of QR in the resistant E. coli isolates, S83L was the most prevalent in gyrase subunit A either alone or combined with D87N and D87Y, and three isolates of QR E. coli isolates revealed a topoisomerase IV subunit mutation harboring S80I. 20% of the isolates displayed efflux activity, as NA showed a considerable difference between its zones of inhibition. Conclusion The high prevalence of antimicrobial-resistant E. coli, with concern for QR strains harboring different resistance mechanisms in poultry meat and beef, threatens the public’s health. Thus, standard manufacturing procedures and adequate hygiene conditions must be followed in all phases of meat preparation, production, and consumption, and public knowledge should be improved.
... E. coli is an extensive and diversified bacterial group. Most E. coli strains are innocuous, but some strains have developed features, such as the ability to produce toxins, that render them hazardous to humans (Garcia et al. 2010;Woyda et al. 2023). ...
Article
Full-text available
Using chicken litter as an organic fertilizer on land is the most common, cheapest and environmentally safest way to manage the latter generated swiftly from the poultry industry. Raw chicken litter has been applied to field soils where various vegetables are cropped to increase yield or productivity. However, the chicken litter frequently come in contact with different environments, such as water, soil, microbes and vegetation. When chickens defecate, their litters, in a few countries, are particularly reused for the next flock, potentially causing cross-contamination. Due to various contact points in the environment, a high probability of bacterial transmission is predicted, which could lead to infection spread in animals and humans. Consumption of contaminated water, food, and meat could lead to the transmission of deadly infections. Microbes in the chicken litter also affect the grazing animals while feeding on fields duly applied with chicken litter as manure. The maximum permissible limits (MPLs) in the chicken litter for land application should not exceed 106-108 CFU/g for Coliform bacteria. Antibiotics are regularly mixed in the diet or drinking water of chicken grown in marketable poultry farms for treating bacterial diseases. Rampant usage of antimicrobials also results in resistant bacteria's survival in animal excreta. Herein, we surveyed the literature to identify the major bacterial genus harboured in the fields applied with chicken manure to increase soil fertility. Our detailed survey identified different bacterial pathogens from chicken litter samples from different investigations. Most studies showed the prevalence of Campylobacter, Salmonella, Enterococcus, E. coli, Bacillus, Comamonas, Proteus and Citrobacter, including many other bacterial species in the chicken litter samples. This article suggested that chicken litter does not meet the standard parameters for direct application as organic fertilizer in the fields. Before being applied to the ground, chicken litter should be treated to lessen the danger of polluting crops or water supplies by reducing the prevalence of harmful bacteria carrying antibiotic-resistance genes.
... A wide variation in the resistance rates of E. coli isolated from poultry carcasses condemned due to colibacillosis was observed, which may be attributed to geographic differences, environmental factors, such as management and climate, and to the specific phylogenetic characteristic of each strain (Woyda et al. 2023). ...
Article
Full-text available
Desirable characteristics of Staphylococcus sp., Streptococcus sp., Bacillus sp., Klebsiella sp., Escherichia coli, and Pseudomonas pseudoalcaligenes isolated from the trachea of healthy turkeys were evaluated as probiotic candidates in the search for new alternatives to solve antimicrobial resistance issues in poultry. In current study phenotypic and genotypic capacity to produce bacteriocin-like substances, efficacy to inhibit the growth of avian pathogens, susceptibility to antimicrobials of bacteria isolated from the respiratory microbiota of healthy turkeys, and the presence of virulence-associated genes (VAGs) predictors of Avian Pathogenic Escherichia coli (APEC) were evaluated. Nine E. coli and one Klebsiella sp. strains produced bacteriocin-like substances, and all harbored the cvaA gene. Some strains also showed antagonistic activity against APEC. Multidrug-resistant profile was found in 54% of the strains. Six strains of bacteriocin-like substances producing E. coli also harbored 3–5 VAGs. The study showed that two bacterial genuses (Klebsiella sp. and E. coli) present desirable probiotic characteristics. Our results identified strains with potential for poultry’s respiratory probiotic.
... Libraries were sequenced on the MiSeq platform with 250-bp paired end reads. Genome assembly, antimicrobial resistance gene identification, virulence factor identification, plasmid replicon identification, phage region identification and genome annotation were done using Reads2Resistome pipeline v0.0.2 (Woyda, Oladeinde and Abdo, 2023). Online ResFinder (Bortolaia et al., 2020) for annotation of acquired resistance genes and additional resistance gene identification was performed using the Resistance Gene Identifier v5.1.1 (RIG) and antibacterial biocide and metal resistance genes were identified with BacMet (Pal et al., 2014). ...
Preprint
Full-text available
Campylobacter infections are a leading cause of bacterial diarrhea in humans globally. Infections are due to consumption of contaminated food products and are highly associated with chicken meat, with chickens being an important reservoir for Campylobacter . Here, we characterized the genetic diversity of Campylobacter species detected in broiler chicken litter over three consecutive flocks and determined their antimicrobial resistance and virulence factor profiles. Antimicrobial susceptibility testing and whole genome sequencing were performed on Campylobacter jejuni (n = 39) and Campylobacter coli (n = 5) isolates. All C. jejuni isolates were susceptible to all antibiotics tested while C. coli (n =4) were resistant to only tetracycline and harbored the tetracycline-resistant ribosomal protection protein (TetO). Virulence factors differed within and across grow houses but were explained by the isolates’ flock cohort, species and multilocus sequence type. Virulence factors involved in the ability to invade and colonize host tissues and evade host defenses were absent from flock cohort 3 C. jejuni isolates as compared to flock 1 and 2 isolates. Our results show that virulence factors and antimicrobial resistance genes differed by the isolates’ multilocus sequence type and by the flock cohort they were present in. These data suggest that the house environment and litter management practices performed imposed selective pressures on antimicrobial resistance genes and virulence factors. In particular, the absence of key virulence factors within the final flock cohort 3 isolates suggests litter reuse selected for Campylobacter strains that are less likely to colonize the chicken host. Importance Campylobacter is a leading cause of foodborne illness in the United States due to the consumption of contaminated food products or from mishandling of food products, often associated with chicken meat. Campylobacter is common in the microbiota of avian and mammalian gut; however, the acquisition of antimicrobial resistance genes and virulence factors may result in strains that pose a significant threat to public health. Although there are studies that have investigated the genetic diversity of Campylobacter strains isolated from post-harvest chicken samples, there is limited data on the genome characteristics of isolates recovered from pre-harvest broiler production. In this study, we show that Campylobacter jejuni and Campylobacter coli that differ in their carriage of antimicrobial resistance and virulence factors may differ in their ability to evade host defense mechanisms and colonize the gut of chickens and humans. Furthermore, we found that differences in virulence factor profiles were explained by the species of Campylobacter and its multilocus sequence type.
Preprint
Full-text available
Background Escherichia coli is commonly used as an indicator for antimicrobial resistance (AMR) in food, animal, environment, and human surveillance systems. Our study aimed to characterize AMR in E. coli isolated from retail meat purchased from grocery stores in North Carolina, USA as part of the National Antimicrobial Resistance Monitoring System (NARMS). Methods Retail chicken (breast, n=96; giblets, n=24), turkey (n=96), and pork (n=96) products were purchased monthly from different counties in North Carolina during 2022. Label claims on packages regarding antibiotic use were recorded at collection. E. coli was isolated from meat samples using culture-based methods and isolates were characterized for antimicrobial resistance using whole genome sequencing. Multi-locus sequence typing, phylogroups, and a single nucleotide polymorphism (SNP)-based maximum-likelihood phylogenic tree were generated. Data were analyzed statistically to determine differences between antibiotic use claims and meat type. Results Of 312 retail meat samples, 138 (44.2%) were positive for E. coli , with turkey (78/138; 56.5%) demonstrating the highest prevalence. Prevalence was lower in chicken (41/138; 29.7%) and pork (19/138;13.8%). Quality sequence data was available from 84.8% (117/138) of the E. coli isolates, which included 72 (61.5%) from turkey, 27 (23.1%) from chicken breast, and 18 (15.4%) from pork. Genes associated with AMR were detected in 77.8% (91/117) of the isolates and 35.9% (42/117) were defined as MDR (≥3 distinct classes of antimicrobials). Commonly observed AMR genes included tetB (35%), tetA (24.8%), aph(3’’)-lb (24.8%), and bla TEM-1 (20.5%), the majority of which originated from turkey isolates. Antibiotics use claims had no statistical effect on MDR E. coli isolates from the different meat types ( X ² =2.21, p =0.33). MDR was observed in isolates from meat products with labels indicating “no claims” (n=29; 69%), “no antibiotics ever” (n=9; 21.4%), and “organic” (n=4; 9.5%). Thirty-four different replicon types were observed. AMR genes were carried on plasmids in 17 E. coli isolates, of which 15 (88.2%) were from turkey and two (11.8%) from chicken. Known sequence types (STs) were described for 81 E. coli isolates, with ST117 (8.5%), ST297 (5.1%), and ST58 (3.4%) being the most prevalent across retail meat types. The most prevalent phylogroups were B1 (29.1%) and A (28.2%). Five clonal patterns were detected among isolates. Conclusions E. coli prevalence and the presence of AMR and MDR were highest in turkey retail meat. The lack of an association between MDR E. coli in retail meat and antibiotic use claim, including those with no indication of antimicrobial use, suggests that additional research is required to understand the origin of resistance. The presence of ST117, an emerging human pathogen, warrants further surveillance. The isolates were distinctly diverse suggesting an instability in population dynamics.
Preprint
Full-text available
Salmonella infections are a leading cause of bacterial food-borne illness worldwide. Infections are highly associated with the consumption of contaminated food, and in particular, chicken meat. Understanding how management practices and environmental factors influence Salmonella populations in broiler chicken production may aid in reducing the risk of food-borne illness in humans. Utilizing whole genome sequencing with antimicrobial and heavy metal resistance, virulence factor and plasmid identification, we have characterized the genetic diversity of Salmonella enterica isolates (n = 55) obtained from broiler chicken litter. S. enterica isolates were recovered from the litter of broiler chickens over three consecutive flocks in four broiler houses on a single integrated farm in Georgia, USA. The chickens were raised under a newly adopted “No Antibiotics Ever” program and copper sulfate was administered via drinking water. In-silico serovar prediction identified three S. enterica serovars: Enteritidis (n = 12), Kentucky (n = 40) and Senftenberg (n = 3). Antimicrobial susceptibility testing revealed that only one S . Kentucky isolate was resistant to streptomycin, while the remaining isolates were susceptible to all antibiotics tested. Metal resistance operons, including copper and silver, were identified chromosomally and on plasmids in serovar Senftenberg and Kentucky isolates, respectively. Serovar Kentucky isolates harboring metal resistance operons were the only Salmonella isolates recovered from the litter of third flock cohort. These results suggest the addition of copper sulfate to drinking water may have selected for S. Kentucky isolates harboring plasmid-borne copper resistance genes and may explain their persistence in litter from flock to flock. Importance Salmonella foodborne illnesses are the leading cause of hospitalizations and deaths, resulting in a high economic burden on the healthcare system. Globally, chicken meat is one of the highest consumed meats and is a predominant source of foodborne illness. The severity of Salmonella infections depends on the presence of antimicrobial resistance genes and virulence factors. While there are many studies which have investigated Salmonella strains isolated from post-harvest chicken samples, there is a gap in our understanding of the prevalence and persistence of Salmonella in pre-harvest and in particular their makeup of antibiotic resistance genes, virulence factors and metal resistance genes. The objective of this study was to determine how on-farm management practices and environmental factors influence Salmonella persistence, as well as the antimicrobial resistance genes and virulence factors they harbor. In this study we demonstrate that broiler chickens raised without antibiotics are less likely to harbor antibiotic resistance, however the practice of adding acidified copper sulfate to drinking water may select for strains carrying metal resistant genes.
Article
Full-text available
VRprofile2 is an updated pipeline that rapidly identifies diverse mobile genetic elements in bacterial genome sequences. Compared with the previous version, three major improvements were made. First, the user-friendly visualization could aid users in investigating the antibiotic resistance gene cassettes in conjunction with various mobile elements in the multiple resistance region with mosaic structure. VRprofile2 could compare the predicted mobile elements to the collected known mobile elements with similar architecture. A new mobilome indicator was proposed to give an overall estimation of the mobilome size in individual bacterial genomes. Second, the relationship between antibiotic resistance genes, mobile elements, and host strains would be efficiently examined with the aid of predicted strain's sequence typing, the incompatibility group and the transferability of plasmids. Finally, the updated back-end database, MobilomeDB2, now collected nearly a thousand active mobile elements retrieved from literature or based on prediction. The pre-computed results of the antibiotic resistance gene-carrying mobile elements of >5500 ESKAPEE genomes were also provided. We expect that VRprofile2 will provide better support for researchers interested in bacterial mobile elements and the dissemination of antibiotic resistance. VRprofile2 is freely available to all users without any login requirement at https://tool2-mml.sjtu.edu.cn/VRprofile.
Article
Full-text available
Fostering a "balanced" gut microbiome through the administration of beneficial microbes that can competitively exclude pathogens has gained a lot of attention and use in human and animal medicine. However, little is known about how microbes affect the horizontal gene transfer of antimicrobial resistance (AMR). To shed more light on this question, we challenged neonatal broiler chicks raised on reused broiler chicken litter-a complex environment made up of decomposing pine shavings, feces, uric acid, feathers, and feed-with Salmonella enterica serovar Heidelberg (S. Heidelberg), a model pathogen. Neonatal chicks challenged with S. Heidelberg and raised on reused litter were more resistant to S. Heidelberg cecal colonization than chicks grown on fresh litter. Furthermore, chicks grown on reused litter were at a lower risk of colonization with S. Heidelberg strains that encoded AMR on IncI1 plasmids. We used 16S rRNA gene sequencing and shotgun metagenomics to show that the major difference between chicks grown on fresh litter and those grown on reused litter was the microbiome harbored in the litter and ceca. The microbiome of reused litter samples was more uniform and enriched in functional pathways related to the biosynthesis of organic and antimi-crobial molecules than that in fresh litter samples. We found that Escherichia coli was the main reservoir of plasmids encoding AMR and that the IncI1 plasmid was maintained at a significantly lower copy per cell in reused litter compared to fresh litter. These findings support the notion that commensal bacteria play an integral role in the horizontal transfer of plasmids encoding AMR to pathogens like Salmonella. IMPORTANCE Antimicrobial resistance spread is a worldwide health challenge, stemming in large part from the ability of microorganisms to share their genetic material through horizontal gene transfer. To address this issue, many countries and international organizations have adopted a One Health approach to curtail the proliferation of antimicrobial-resistant bacteria. This includes the removal and reduction of antibiotics used in food animal production and the development of alternatives to antibiotics. However, there is still a significant knowledge gap in our understanding of how resistance spreads in the absence of antibiotic selection and the role commen-sal bacteria play in reducing antibiotic resistance transfer. In this study, we show that commensal bacteria play a key role in reducing the horizontal gene transfer of antibiotic resistance to Salmonella, provide the identity of the bacterial species that potentially perform this function in broiler chickens, and also postulate the mechanism involved.
Article
Full-text available
Escherichia coli is a priority foodborne pathogen of public health concern and phenotypic serotyping provides critical information for surveillance and outbreak detection activities. Public health and food safety laboratories are increasingly adopting whole-genome sequencing (WGS) for characterizing pathogens, but it is imperative to maintain serotype designations in order to minimize disruptions to existing public health workflows. Multiple in silico tools have been developed for predicting serotypes from WGS data, including SRST2, SerotypeFinder and EToKi EBEis, but these tools were not designed with the specific requirements of diagnostic laboratories, which include: speciation, input data flexibility (fasta/fastq), quality control information and easily interpretable results. To address these specific requirements, we developed ECTyper ( https://github.com/phac-nml/ecoli_serotyping ) for performing both speciation within Escherichia and Shigella , and in silico serotype prediction. We compared the serotype prediction performance of each tool on a newly sequenced panel of 185 isolates with confirmed phenotypic serotype information. We found that all tools were highly concordant, with 92–97 % for O-antigens and 98–100 % for H-antigens, and ECTyper having the highest rate of concordance. We extended the benchmarking to a large panel of 6954 publicly available E. coli genomes to assess the performance of the tools on a more diverse dataset. On the public data, there was a considerable drop in concordance, with 75–91 % for O-antigens and 62–90 % for H-antigens, and ECTyper and SerotypeFinder being the most concordant. This study highlights that in silico predictions show high concordance with phenotypic serotyping results, but there are notable differences in tool performance. ECTyper provides highly accurate and sensitive in silico serotype predictions, in addition to speciation, and is designed to be easily incorporated into bioinformatic workflows.
Article
Full-text available
Salmonella enterica is common foodborne pathogen that generates both enteric and systemic infections in hosts. Antibiotic resistance is common is certain serovars of the pathogen and of great concern to public health. Recent reports have documented the co-occurrence of metal resistance with antibiotic resistance in one serovar of S. enterica. Therefore, we sought to identify possible co-occurrence in a large genomic dataset. Genome assemblies of 56,348 strains of S. enterica comprising 20 major serovars were downloaded from NCBI. The downloaded assemblies were quality controlled and in silico serotyped to ensure consistency and avoid improper annotation from public databases. Metal and antibiotic resistance genes were identified in the genomes as well as plasmid replicons. Co-occurrent genes were identified by constructing a co-occurrence matrix and grouping said matrix using k-means clustering. Three groups of co-occurrent genes were identified using k-means clustering. Group 1 was comprised of the pco and sil operons that confer resistance to copper and silver, respectively. Group 1 was distributed across four serovars. Group 2 contained the majority of the genes and little to no co-occurrence was observed. Metal and antibiotic co-occurrence was identified in group 3 that contained genes conferring resistance to: arsenic, mercury, beta-lactams, sulfonamides, and tetracyclines. Group 3 genes were also associated with an IncQ1 class plasmid replicon. Metal and antibiotic co-occurrence from group 3 genes is mostly isolated to one clade of S. enterica I 4,[5],12:i:-
Article
Full-text available
Simple Summary: This revision is about the problem of Escherichia coli as a commensal and pathogenic bacterium among food-producing animals and health implications. Escherichia coli may play an important ecological role and can be used as a bioindicator of antimicrobial resistance. All animal species used for food production, as well as humans, carry E. coli in their intestinal tract; plus, the genetic flexibility and adaptability of this bacteria to constantly changing environments allows it to acquire a great number of antimicrobial resistance mechanisms. The majority of E. coli strains are commensals inhabiting the intestinal tract of humans and warm-blooded animals and rarely causes diseases. However, E. coli also remains as one of the most frequent causes of several common bacterial infections in humans and animals. All over the word, antibiotic resistance is commonly detected among commensal bacteria from food-producing animals, raising important questions on the potential impact of antibiotic use in animals and the possible transmission of these resistant bacteria to humans through the food chain. The use, in food-producing animals, of antibiotics that are critically important in human medicine has been implicated in the emergence of new forms of resistant bacteria, including new strains of multidrug-resistant foodborne bacteria, such as extended spectrum β-lactamase (ESBL)-producing E. coli.
Article
Full-text available
Objectives: WGS-based antimicrobial susceptibility testing (AST) is as reliable as phenotypic AST for several antimicrobial/bacterial species combinations. However, routine use of WGS-based AST is hindered by the need for bioinformatics skills and knowledge of antimicrobial resistance (AMR) determinants to operate the vast majority of tools developed to date. By leveraging on ResFinder and PointFinder, two freely accessible tools that can also assist users without bioinformatics skills, we aimed at increasing their speed and providing an easily interpretable antibiogram as output. Methods: The ResFinder code was re-written to process raw reads and use Kmer-based alignment. The existing ResFinder and PointFinder databases were revised and expanded. Additional databases were developed including a genotype-to-phenotype key associating each AMR determinant with a phenotype at the antimicrobial compound level, and species-specific panels for in silico antibiograms. ResFinder 4.0 was validated using Escherichia coli (n = 584), Salmonella spp. (n = 1081), Campylobacter jejuni (n = 239), Enterococcus faecium (n = 106), Enterococcus faecalis (n = 50) and Staphylococcus aureus (n = 163) exhibiting different AST profiles, and from different human and animal sources and geographical origins. Results: Genotype-phenotype concordance was ≥95% for 46/51 and 25/32 of the antimicrobial/species combinations evaluated for Gram-negative and Gram-positive bacteria, respectively. When genotype-phenotype concordance was <95%, discrepancies were mainly linked to criteria for interpretation of phenotypic tests and suboptimal sequence quality, and not to ResFinder 4.0 performance. Conclusions: WGS-based AST using ResFinder 4.0 provides in silico antibiograms as reliable as those obtained by phenotypic AST at least for the bacterial species/antimicrobial agents of major public health relevance considered.
Article
Full-text available
Widespread use of antibiotics has enhanced the evolution of highly resilient pathogens and poses a severe risk to human health via coselection of antibiotic resistance genes (ARGs) and virulence factors (VFs). In this study, we rigorously evaluate the abundance relationship and physical linkage between ARGs and VFs by performing a comprehensive analysis of 9,070 bacterial genomes isolated from multiple species and hosts. The coexistence of ARGs and VFs was observed in bacteria across distinct phyla, pathogenicities, and habitats, especially among human-associated pathogens. The coexistence patterns of gene elements in different habitats and pathogenicity groups were similar, presumably due to frequent gene transfer. A shorter intergenic distance between mobile genetic elements and ARGs/VFs was detected in human/animal-associated bacteria, indicating a higher transfer potential. Increased accumulation of exogenous ARGs/VFs in human pathogens highlights the importance of gene acquisition in the evolution of human commensal bacteria. Overall, the findings provide insights into the genic features of combinations of ARG-VF and expand our understanding of ARG-VF coexistence in bacteria. IMPORTANCE Antibiotic resistance has become a serious global health concern. Despite numerous case studies, a comprehensive analysis of ARG and VF coexistence in bacteria is lacking. In this study, we explore the coexistence profiles of ARGs and VFs in diverse categories of bacteria by using a high-resolution bioinformatics approach. We also provide compelling evidence of unique ARG-VF gene pairs coexisting in specific bacterial genomes and reveal the potential risk associated with the coexistence of ARGs and VFs in organisms in both clinical settings and environments.
Preprint
Full-text available
The bacterial resistome is the collection of all the antibiotic resistance genes, virulence genes, and other resistance elements within a bacterial isolate genome including plasmids and bacteriophage regions. Accurately characterizing the resistome is crucial for prevention and mitigation of emerging antibiotic resistance threats to animal and human health. Reads2Resistome is a tool which allows researchers to assemble and annotate bacterial genomes using long or short read sequencing technologies or both in a hybrid approach. Using a massively parallel analysis pipeline, Reads2Resistome performs assembly, annotation and resistome characterization with the goal of producing an accurate and comprehensive description of a bacterial genome and resistome contents. Key features of the Reads2Resistome pipeline include quality control of input sequencing reads, genome assembly, genome annotation, resistome characterization and alignment. All prerequisite dependencies come packaged together in a single suit which can easily be downloaded and run on Linux and Mac operating systems. Availability Reads2Resistome is freely available as an open-source package under the MIT license, and can be downloaded via GitHub ( https://github.com/BioRRW/Reads2Resistome ).
Article
Ggtree is an R/Bioconductor package for visualizing tree‐like structures and associated data. After 5 years of continual development, ggtree has been evolved as a package suite that contains treeio for tree data input and output, tidytree for tree data manipulation, and ggtree for tree data visualization. Ggtree was originally designed to work with phylogenetic trees, and has been expanded to support other tree‐like structures, which extends the application of ggtree to present tree data in other disciplines. This article contains five basic protocols describing how to visualize trees using the grammar of graphics syntax, how to visualize hierarchical clustering results with associated data, how to estimate bootstrap values and visualize the values on the tree, how to estimate continuous and discrete ancestral traits and visualize ancestral states on the tree, and how to visualize a multiple sequence alignment with a phylogenetic tree. The ggtree package is freely available at https://www.bioconductor.org/packages/ggtree . © 2020 by John Wiley & Sons, Inc. Basic Protocol 1 : Using grammar of graphics for visualizing trees Basic Protocol 2 : Visualizing hierarchical clustering using ggtree Basic Protocol 3 : Visualizing bootstrap values as symbolic points Basic Protocol 4 : Visualizing ancestral status Basic Protocol 5 : Visualizing a multiple sequence alignment with a phylogenetic tree