12
Genomic Diversity, Virulence, and Antimicrobial Resistance of Klebsiella pneumoniae Strains from Cows and Humans Yongqiang Yang, a,b,c Catherine H. Higgins, a Ibraheem Rehman, a Klibs N. Galvao, d Ilana L. Brito, e Marcela L. Bicalho, a Jeongmin Song, f Hongning Wang, b Rodrigo C. Bicalho a a Department of Population Medicine and Diagnostic Sciences, College of Veterinary Medicine, Cornell University, Ithaca, New York, USA b Animal Disease Prevention and Food Safety Key Laboratory of Sichuan Province, College of Life Science, Sichuan University, Chengdu, China c School of Pharmaceutical Sciences (Shenzhen), Sun Yat‐sen University, Guangzhou, China d Department of Large Animal Clinical Sciences, College of Veterinary Medicine, University of Florida, Gainesville, Florida, USA e Nancy E. and Peter C. Meinig School of Biomedical Engineering, College of Engineering, Cornell University, Ithaca, New York, USA f Department of Microbiology and Immunology, College of Veterinary Medicine, Cornell University, Ithaca, New York, USA ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections are rapidly becoming untreatable due to the emergence of multidrug-resistant (MDR) strains. However, little is known about the relationship between bovine and human K. pneumoniae isolates at the genome pop- ulation level. Here, we investigated the genomic structures, pangenomic profiles, vir- ulence determinants, and resistomes of 308 K. pneumoniae isolates from humans and dairy cows, including 96 newly sequenced cow isolates. We identified 177 functional protein families that were significantly different across human and bovine isolates; genes expressing proteins related to metal ion (iron, zinc, and calcium) metabolism were significantly more prevalent among the bovine isolates. Siderophore systems were found to be prevalent in both the bovine and the human isolates. In addition, we found that the Klebsiella ferric uptake operon kfuABC was significantly more prevalent in clinical mastitis cases than in healthy cows. Furthermore, on two dairy farms, we identified a unique IncN-type plasmid, pC5, coharboring bla CTX-M-1 and mph(A) genes, which confer resistance to cephalosporins and macrolides, respec- tively. We provide here the complete annotated sequence of this plasmid. IMPORTANCE We demonstrate here the genetic diversity of K. pneumoniae isolates from dairy cows and the mixed phylogenetic lineages between bovine and human isolates. The ferric uptake operon kfuABC genes were more prevalent in strains from clinical mastitis cows. Furthermore, we report the emergence of an IncN-type plas- mid carrying the bla CTX-M-1 and mph(A) genes among dairy farms in the United States. Our study evaluated the genomic diversity of the bovine and human isolates, and the findings uncovered different profiles of virulence determinants among bo- vine and human K. pneumoniae isolates at the genome population level. KEYWORDS antimicrobial resistance, genomic epidemiology, Klebsiella pneumoniae, virulence K lebsiella pneumoniae is a Gram-negative, rod-shaped, encapsulated, facultative anaerobe that is commonly found in the mouth, skin, and intestines of humans and animals. As an opportunistic pathogen in humans, K. pneumoniae primarily attacks immunocompromised individuals and has become a leading cause of community- acquired and hospital-acquired infections in humans (1). K. pneumoniae is a clinically important species and causes serious nosocomial infections, such as septicemia, pneu- monia, urinary tract infection, surgical site infection, and soft tissue infection (2). Nosocomial infections caused by K. pneumoniae are a leading cause of morbidity and Citation Yang Y, Higgins CH, Rehman I, Galvao KN, Brito IL, Bicalho ML, Song J, Wang H, Bicalho RC. 2019. Genomic diversity, virulence, and antimicrobial resistance of Klebsiella pneumoniae strains from cows and humans. Appl Environ Microbiol 85:e02654-18. https:// doi.org/10.1128/AEM.02654-18. Editor Christopher A. Elkins, Centers for Disease Control and Prevention Copyright © 2019 American Society for Microbiology. All Rights Reserved. Address correspondence to Rodrigo C. Bicalho, [email protected]. Received 2 November 2018 Accepted 21 December 2018 Accepted manuscript posted online 4 January 2019 Published EVOLUTIONARY AND GENOMIC MICROBIOLOGY crossm March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 1 Applied and Environmental Microbiology 6 March 2019 on February 20, 2021 by guest http://aem.asm.org/ Downloaded from

Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

Genomic Diversity, Virulence, and Antimicrobial Resistance ofKlebsiella pneumoniae Strains from Cows and Humans

Yongqiang Yang,a,b,c Catherine H. Higgins,a Ibraheem Rehman,a Klibs N. Galvao,d Ilana L. Brito,e Marcela L. Bicalho,a

Jeongmin Song,f Hongning Wang,b Rodrigo C. Bicalhoa

aDepartment of Population Medicine and Diagnostic Sciences, College of Veterinary Medicine, Cornell University, Ithaca, New York, USAbAnimal Disease Prevention and Food Safety Key Laboratory of Sichuan Province, College of Life Science, Sichuan University, Chengdu, ChinacSchool of Pharmaceutical Sciences (Shenzhen), Sun Yat‐sen University, Guangzhou, ChinadDepartment of Large Animal Clinical Sciences, College of Veterinary Medicine, University of Florida, Gainesville, Florida, USAeNancy E. and Peter C. Meinig School of Biomedical Engineering, College of Engineering, Cornell University, Ithaca, New York, USAfDepartment of Microbiology and Immunology, College of Veterinary Medicine, Cornell University, Ithaca, New York, USA

ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humansand dairy cows, and these infections are rapidly becoming untreatable due to theemergence of multidrug-resistant (MDR) strains. However, little is known about therelationship between bovine and human K. pneumoniae isolates at the genome pop-ulation level. Here, we investigated the genomic structures, pangenomic profiles, vir-ulence determinants, and resistomes of 308 K. pneumoniae isolates from humans anddairy cows, including 96 newly sequenced cow isolates. We identified 177 functionalprotein families that were significantly different across human and bovine isolates;genes expressing proteins related to metal ion (iron, zinc, and calcium) metabolismwere significantly more prevalent among the bovine isolates. Siderophore systemswere found to be prevalent in both the bovine and the human isolates. In addition,we found that the Klebsiella ferric uptake operon kfuABC was significantly moreprevalent in clinical mastitis cases than in healthy cows. Furthermore, on two dairyfarms, we identified a unique IncN-type plasmid, pC5, coharboring blaCTX-M-1 andmph(A) genes, which confer resistance to cephalosporins and macrolides, respec-tively. We provide here the complete annotated sequence of this plasmid.

IMPORTANCE We demonstrate here the genetic diversity of K. pneumoniae isolatesfrom dairy cows and the mixed phylogenetic lineages between bovine and humanisolates. The ferric uptake operon kfuABC genes were more prevalent in strains fromclinical mastitis cows. Furthermore, we report the emergence of an IncN-type plas-mid carrying the blaCTX-M-1 and mph(A) genes among dairy farms in the UnitedStates. Our study evaluated the genomic diversity of the bovine and human isolates,and the findings uncovered different profiles of virulence determinants among bo-vine and human K. pneumoniae isolates at the genome population level.

KEYWORDS antimicrobial resistance, genomic epidemiology, Klebsiella pneumoniae,virulence

Klebsiella pneumoniae is a Gram-negative, rod-shaped, encapsulated, facultativeanaerobe that is commonly found in the mouth, skin, and intestines of humans and

animals. As an opportunistic pathogen in humans, K. pneumoniae primarily attacksimmunocompromised individuals and has become a leading cause of community-acquired and hospital-acquired infections in humans (1). K. pneumoniae is a clinicallyimportant species and causes serious nosocomial infections, such as septicemia, pneu-monia, urinary tract infection, surgical site infection, and soft tissue infection (2).Nosocomial infections caused by K. pneumoniae are a leading cause of morbidity and

Citation Yang Y, Higgins CH, Rehman I, GalvaoKN, Brito IL, Bicalho ML, Song J, Wang H,Bicalho RC. 2019. Genomic diversity, virulence,and antimicrobial resistance of Klebsiellapneumoniae strains from cows and humans.Appl Environ Microbiol 85:e02654-18. https://doi.org/10.1128/AEM.02654-18.

Editor Christopher A. Elkins, Centers forDisease Control and Prevention

Copyright © 2019 American Society forMicrobiology. All Rights Reserved.

Address correspondence to Rodrigo C. Bicalho,[email protected].

Received 2 November 2018Accepted 21 December 2018

Accepted manuscript posted online 4January 2019Published

EVOLUTIONARY AND GENOMIC MICROBIOLOGY

crossm

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 1Applied and Environmental Microbiology

6 March 2019

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 2: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

mortality in immunocompromised people, with fatality rates being upwards of 50%despite the use of appropriate antibiotic therapy (3, 4).

In veterinary medicine, K. pneumoniae is of interest because it commonly causesclinical mastitis in dairy herds. Mastitis is a highly prevalent disease in dairy cows andcreates an incontestable economic burden in the dairy industry worldwide, in additionto being a serious animal welfare issue because of the pain associated with theinfection (5). The economic losses associated with mastitis are predominantly due todecreased milk production, discarded milk, treatment costs, and premature culling (6, 7).Coliform organisms are among the highest-incidence organisms on farms with otherwiseexcellent udder health management. The implementation of programs for mastitis controlhas reduced the prevalence of important contagious pathogens, and currently, approxi-mately 40% of all clinical mastitis cases are associated with opportunistic Gram-negativebacteria, such as Klebsiella spp., Escherichia coli, Pseudomonas spp., and Pasteurella spp. (8,9, 10). Among the major mastitis pathogens causing mastitis, coliform organisms such asKlebsiella spp. and E. coli cause the most severe clinical cases, with greater milk loss and anaverage cost of more than $221.03 per case (11). Studies have shown differences in thepathogenicity of Klebsiella spp. as a mastitis pathogen and other Gram-negative organisms(12, 13). Klebsiella spp. cause longer intramammary infections than E. coli (14), more severeclinical episodes than Serratia spp. and E. coli (15, 16), and a greater milk production lossand risk of culling than intramammary infections caused by E. coli (17). Therefore, due to theseverity of bovine clinical mastitis, the low effectiveness of antibiotic treatments, and thelack of advancement in preventive measures against Klebsiella infection, K. pneumoniae is aparticularly concerning pathogen.

Whole-genome analysis is the ideal approach to building a robust phylogeny in recentlyemerged pathogens to explore their diverse backgrounds through the identification ofsingle nucleotide polymorphisms (SNPs) or other genetic variants. A recent pangenomicanalysis of K. pneumoniae isolates from humans and bovines found that genes namedregulators of mucoid phenotype A (rmpA) and A2 (rmpA2) as well as a cluster of sidero-phore genes (kfuABC and kvgAS) are associated with invasive infection in humans (18).However, the analysis of the genetic determinants of the bovine isolates was limited, andthe relatedness of the human and bovine K. pneumoniae isolates was not explored. A morecomprehensive study with a larger number of isolates from cows would allow for explo-ration of the relatedness of human isolates and bovine isolates.

Here, we assembled a collection of K. pneumoniae strains recovered from 143mastitic cows with various degrees of disease severity and sequenced the genomes of96 distinct isolates to compare their genomic structures, virulence factors, and antimi-crobial resistance (AMR) genes with those of the K. pneumoniae isolates studied by Holtet al. (18). We investigated the pangenomic gene functions of 308 isolates in total andidentified a group of virulence genes that had a different distribution between K.pneumoniae isolates collected from dairy cows and humans. Furthermore, we found aresistance plasmid not previously identified in the United States and clarified its geneticcontext according to the closed plasmid structure.

RESULTSDiversity and AMR profile of K. pneumoniae on dairy farms. We isolated 143

nonreplicate K. pneumoniae strains (see Fig. S1 and Data Set S1 in the supplementalmaterial) from mastitic dairy cows from four farms located in upstate New York.Molecular typing of the isolates using pulsed-field gel electrophoresis (PFGE) identified97 distinct PFGE groups, revealing a high genetic diversity within the isolate collection.Capsule locus typing of those 143 isolates showed that 127 (88.8%) were assigned to44 known capsule loci (Fig. S1). The capsule loci KN3 (30/143, 21%), K13 (15/143, 11%),and KN1 (10/143, 7%) were the most prevalent in our sample set, accounting forapproximately 39% of all infections. High genetic diversity was also reflected by thesequence types (STs) of those strains. Among the 96 newly sequenced strains, 46possessed new alleles of the housekeeping genes used for multilocus sequence typing

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 2

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 3: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

(MLST), and no predominant STs were found. These new alleles were submitted to theMLST database and were assigned 43 new STs.

Overall, the AMR profiles were diverse among these strains, and 40% (57/143) wereresistant to one or more antimicrobial agents (Fig. S1; Data Set S1). The agent to whichthe highest prevalence of resistance was detected was streptomycin (29.4%, 42/143),followed by tetracycline (5.6%, 8/143) and gentamicin (4.2%, 6/143). Using the genomicdata for the 96 newly sequenced isolates from mastitic cows, we detected 17 AMR geneswith various prevalences (Fig. S2). The fosA (fosfomycin resistance), oqxAB (quinolonesresistance), blaAmpH (�-lactamase), and blaSHV (�-lactamase) genes were common resis-tance genes found in our isolate collection. The resistance to streptomycin in our isolatecollection was attributed to the strA and strB genes, located on an IncHI1B-type plasmid in27 of the 96 strains. Analysis of the genetic environment revealed that the strA and strBgenes were flanked by transposon Tn5393, the most common mobile element thatmediates the transmission of strAB genes across species (19).

Identification of an IncN-type plasmid carrying blaCTX-M-1 and mph(A). Wefound that four isolates were resistant to ceftiofur, with the MIC being 512 �g/ml.Among those, three strains that had genomic sequences were found to coharbor theblaCTX-M-1 (�-lactamase) and mph(A) (macrolide resistance) genes, which were locatedin the same contig in each strain. Comparison of the sequences by BLASTn analysisshowed that all three contigs shared a high degree of similarity with the sequence ofpL2-43 (GenBank accession no. KJ484641), an IncN-type plasmid found in an E. coliisolate recovered from a lamb in Switzerland in 2014 (20). Analysis of the plasmidsequences confirmed that they carried the same plasmid (named pC5) that coharboredthe blaCTX-M-1 and mph(A) genes. The complete sequence of pC5 (GenBank accessionno. MF953243) is 41,608 bp in size and carries 51 open reading frames (ORFs) (Fig. 1;Table S1). The blaCTX-M-1 gene is flanked by an ISEcp1 element in its upstream region,which is disrupted by an intact IS26 element (820 bp) in the opposite orientation. Thesequences between the blaCTX-M-1 and mph(A) genes (1,572 bp) contain two ORFs: oneORF carries a truncated mrx gene (1,272 bp), and the other ORF is similar to ORF477 butis truncated by an inverted right repeat (IRR) of the ISEcp1 element (23 bp). The mph(A)gene was mediated by another IS26 element with the same orientation. We then

FIG 1 Plasmid structures of pC5 and reference plasmids. (A) Circular genetic map of plasmid pC5 (GenBank accession no. MF953243) and reference plasmidspL2-43 (GenBank accession no. KJ484641), pKC394 (GenBank accession no. HM138652), pH1038-142 (GenBank accession no. KJ484634), pVI (GenBank accessionno. LT795508), pHHA45 (GenBank accession no. JX065630), pQNR2078 (GenBank accession no. HE613857), and pKC396 (GenBank accession no. HM138653). Themap was drawn using the BLAST Ring Image Generator (BRIG) program (http://sourceforge.net/projects/brig/). (B) Schematic presentation of major structuralfeatures of pC5 in comparison with the reference plasmids pL2-43, pKC394, pH1038-142, pVI, pHHA45, pQNR2078, and pKC396. Areas shaded in gray indicatehomologous regions of �99% nucleotide sequence identity in the plasmid scaffold regions. ORFs are portrayed by arrows to indicate the direction oftranscription and colored on the basis of their predicted gene functions. Antimicrobial resistance genes are indicated by red boxes. Blue boxes denote mobilegenetic elements. Green boxes indicate ORFs encoding hypothetical proteins. The figure is drawn to scale.

Klebsiella pneumoniae from Cows and Humans Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 3

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 4: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

compared the genetic background of pC5 with that of seven publicly available plasmidsequences with high similarity (coverage, �90%; identity, �99%), all of which wereidentified in E. coli strains in European countries (20, 21, 22, 23) (Table 1). They allmaintained a typical IncN plasmid backbone scaffold, including a replicon and genesinvolved in plasmid stability, mutagenesis enhancement, antirestriction functions, andconjugative transfer (24). The difference among these plasmids was in the mosaicregion that carried resistance genes and in the mobile genetic elements represented(Fig. 1). Plasmid pC5 shared an identical genetic neighborhood of the blaCTX-M-1 andmph(A) genes with pHHA45, pKC394, and pVI. However, in plasmid pL2-43, the reverseposition of mph(A)-IS26 was the main structural difference from our plasmid.

Phylogeny of K. pneumoniae from cows and humans. We compared 308 ge-nomes from K. pneumoniae KpI isolates, including 123 isolates from dairy cows and 185isolates from humans. We identified a total of 92,409 core genome SNPs among these308 genomes and built a maximum likelihood (ML) phylogenetic tree based on theseSNPs (Fig. 2). The results revealed a deep branching and scattered population structurethat was broadly classified into 169 distinct phylogenetic lineages. The 123 bovineisolates had diverse population structures and did not form specific clusters. Further-more, the population structure of the 123 bovine isolates was highly mixed with the185 human isolates, and there were 27 lineages containing both human and bovineisolates. We assessed the clonality and clades of 294 isolates with assigned STs from the308 strains as minimum-spanning trees (Fig. 3). Of these, we identified 190 distinct STs.The most prevalent STs were ST23 (n � 10 isolates), ST43 (n � 8), ST11 (n � 7), ST107(n � 7), ST15 (n � 6), ST199 (n � 6), ST14 (n � 5), ST20 (n � 5), and ST48 (n � 5).Furthermore, ST34, ST35, ST37, ST43, ST65, ST107, ST133, ST290, ST294, ST309, andST791 were found in both human and bovine isolates.

Pangenomic analysis identified distinct proteins between cow and humanisolates. We then investigated the functions of all genes across the 308 human andbovine isolates. In total, there were 1,705,306 open reading frames (ORFs) in the 308genomes, with an average of 5,536 ORFs in each genome. All the ORFs could beclassified into 68,420 clusters by sequence identity, which included 24,753 clustersharboring multiple sequences and 43,667 clusters with singleton genes. Functionannotation against the Kyoto Encyclopedia of Genes and Genomes (KEGG) databaseshowed that 72% of the representative genes were assigned to 9,407 distinct KEGG hitswith known functions among different functional categories (Fig. S3; Table S2). Thesefunctions included environmental information processing, genetic information process-ing, cellular processes, carbohydrate metabolism, amino acid metabolism, metabolismof cofactors and vitamins, energy metabolism, human disease-related functions, en-zyme families, and nucleotide metabolism.

We then calculated the distribution of these KEGG assignments. There were 1,800and 2,392 functional units found in all bovine isolates and human isolates, respectively.Among them, 2,684 functional units were found at a prevalence of �95% in bothbovine and human isolates. Nonetheless, there were 177 functional units with signifi-cant differences between the human and bovine isolates (P � 0.0001, �2 test) (Fig. S4;Table S3). Proteins related to metal ion (iron, zinc, and calcium) metabolism weresignificantly more prevalent in the bovine isolates; examples included the Fe3� dicitratetransport protein FecA (81.3% versus 43.2%), the zinc protease PqqL (65.9% versus 7%),and the calcium permeable stress-gated cation channel protein TMEM63 (60.2% versus44.3%). However, the reverse situation was true for proteins related to heavy metalion-related transport; for example, 29.7% of the human isolates but only 2.4% of thebovine isolates carried the mercuric ion transport protein MerT.

Extensive diversity of virulence genes among bovine and human isolates. Intotal, we detected 173 virulence genes in the 308 isolates (Table S4). These included135 and 170 virulence genes found in the bovine and human isolates, respectively, ofwhich 132 genes were common to both types of isolates (Fig. 4). The enterobactin loci,adhesion-related gene clusters, secretion system-related gene clusters, and fimbria

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 4

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 5: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

TAB

LE1

Sum

mar

yof

the

feat

ures

ofth

ene

wly

sequ

ence

dp

C5

and

seve

nre

late

dp

lasm

ids

inth

elit

erat

ure

Isol

ate

Stra

inY

rof

isol

atio

nC

oun

try

Ori

gin

MLS

TPl

asm

idG

enB

ank

acce

ssio

nn

o.Pl

asm

idsi

ze(b

p)

Inc

typ

eRe

sist

ance

gen

e(s)

Cov

erag

ea

(%)

Iden

tity

a

(%)

Sour

ceor

refe

ren

ce

C5

K.pn

eum

onia

e20

16U

SAC

ow27

48p

C5

MF9

5324

341

,608

IncN

bla C

TX

-M-1

,mph

(A)

—b

—Th

isst

udy

L-2

E.co

li—

Switz

erla

ndLa

mb

295

pL2

-43

KJ48

4641

43,2

65In

cNbl

a CT

X-M

-1,m

ph(A

)10

099

2039

4E.

coli

2006

Ger

man

yH

uman

131

pKC

394

HM

1386

5253

,207

IncN

bla C

TX

-M-1

,mph

(A)

9999

21H

-103

8E.

coli

2010

Switz

erla

ndH

uman

New

pH

1038

-142

KJ48

4634

142,

875

IncF

-IncN

bla C

TX

-M-1

,bla

TE

M-1

,tet

R-te

t(A

),st

rAB,

catA

1,su

l3,a

adA

1,df

rA1

9799

20

KV7

E.co

li—

UK

Pig

—p

VILT

7955

0839

,744

IncN

bla C

TX

-M-1

,mph

(A)

9399

Gen

Bank

c

HH

A45

E.co

li20

06D

enm

ark

Pig

—p

HH

A45

JX06

5630

39,5

10In

cNbl

a CT

X-M

-1,m

ph(A

)93

9922

2078

E.co

li—

Ger

man

yH

orse

86p

QN

R207

8H

E613

857

42,3

79In

cNqn

rB19

9199

2339

6E.

coli

2006

Ger

man

yH

uman

131

pKC

396

HM

1386

5344

,216

IncN

bla C

TX

-M-6

591

9921

ap

C5

was

the

quer

yp

lasm

idus

edto

inve

stig

ate

the

sequ

ence

cove

rage

and

iden

tity.

b—

,fea

ture

not

iden

tified

.c G

enBa

nkdi

rect

sub

mis

sion

.

Klebsiella pneumoniae from Cows and Humans Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 5

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 6: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

gene clusters were found in all isolates. Of note, the average number of virulence genesper isolate was significantly higher for the human isolates than for the bovine isolates(70 versus 62; P � 0.0001, Kruskal-Wallis test) (Fig. 4). The virulence genes associatedwith colibactin, aerobactin, allantoinase, yersiniabactin, microcin, salmochelin, and thermpA and rmpA2 genes were rare (0.5% to 5%) among the bovine isolates, whereas theywere found at increased frequency in the human isolates (9% to 35%) (Fig. 4).Furthermore, the combination of colibactin, aerobactin, allantoinase, yersiniabactin,microcin, salmochelin, and the rmpA and rmpA2 genes was frequent in K. pneumoniaeisolates from humans. Among all 123 bovine isolates, we also investigated the distri-bution of virulence genes in healthy cows and mastitic cows. All the virulence genesfound in the healthy group were present in either the subclinical mastitis or clinicalmastitis cases. However, 26 virulence genes were exclusive to the subclinical mastitiscases and 6 were exclusive to the clinical mastitis cases (Fig. 4). Notably, we found thatthe Klebsiella ferric uptake operon kfuABC was significantly more prevalent (P � 0.05, �2

test) in clinical mastitis cases than in subclinical mastitis cases and healthy cows (39%,20%, and 11%, respectively; P � 0.0336, �2 test).

DISCUSSION

The emergence and transmission of multidrug-resistant and virulent K. pneumoniaestrains can cause severe, untreatable infections in humans and animals. However, dueto insufficient data at the population level, the diversity of AMR genes of K. pneumoniaein cows is still unknown. In this study, we identified an IncN-type plasmid, pC5,coharboring the extended-spectrum �-lactamase (ESBL) gene blaCTX-M-1 and the mac-rolide phosphotransferase gene mph(A) in four distinct isolates on two farms. Ceftiofuris one of the most commonly used antibiotics in both adult and young dairy cattle in the

FIG 2 Phylogeny of core genome SNPs in 308 genomes of K. pneumoniae KpI isolates from humans and dairy cows. Thecore genome SNPs were identified by mapping sequencing reads against a reference genome (K. pneumoniae strainNTUH-K2044, GenBank accession no. NC_012731.1). The RAxML program was used to calculate the phylogenetic tree toconstruct a maximum likelihood phylogeny.

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 6

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 7: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

United States, whereas macrolides are mostly used in young dairy cattle (�20 months ofage) for the treatment of respiratory disease (https://www.aphis.usda.gov/animal_health/nahms/dairy/downloads/dairy07/Dairy07_is_AntibioticUse.pdf). The use of ceftiofurand/or macrolides may promote the dissemination of pC5-like plasmids. Seven plas-

FIG 3 Minimum-spanning trees of 294 K. pneumoniae isolates from humans and dairy cows by MLST type. The 294 isolates that had assigned STs were derivedfrom 308 K. pneumoniae isolates from humans and dairy cows. Each node within the tree represents a single ST. The size of the nodes is proportional to thenumber of K. pneumoniae isolates belonging to that ST. The length of the branch between each node is proportional to the number of distinct alleles of sevenhousekeeping genes that differ between the two linked nodes. The 30 STs with the highest numbers of isolates are labeled with different colors.

Klebsiella pneumoniae from Cows and Humans Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 7

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 8: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

mids that had been previously identified in E. coli either in humans or in animals inEuropean countries shared a high similarity with pC5. In the United States, CTX-M-like-producing Enterobacteriaceae strains were first reported in 2003 (25). CTX-M-15 is themost common CTX-M enzyme found in Enterobacteriaceae in humans in the UnitedStates (26), whereas CTX-M-1 is rare. A previous study found 1 CTX-M-1-producingisolate among 208 ESBL-producing K. pneumoniae isolates from humans and 1 CTX-M-1-producing isolate among 163 ESBL-producing E. coli isolates from humans (27). Thewide geographical and source distribution of pC5-like plasmids in European countriesindicated that such IncN plasmids are responsible for the dissemination of blaCTX-M-1

and other resistance genes. Our study reports the emergence of this plasmid in theUnited States.

Our study sequenced 96 strains which belonged to different PFGE groups, and 46 ofthese were new STs. This indicates that there is no dominant strain responsible forinfections with K. pneumoniae in cows. We built a representative pool of genomes frombovine and human strains comprising more than 300 genomes. Both the human andbovine isolates shared a large number of functional modules to form a core genomethat may contribute to the maintenance of bacterial life. However, environmentalchanges and selection pressures have led to differences in some specific functionsamong the human and bovine isolates. For example, we found that proteins involvedin metal ion (iron, zinc, and calcium) metabolism were significantly more prevalent in

FIG 4 Virulence genes in 308 K. pneumoniae KpI isolates. (A) Number of virulence genes per isolate. Box plots indicate the average number of virulence genesin 123 bovine strains and 185 human strains. The means are shown as horizontal bars, with values of 62 and 70 for the bovine and human isolates, respectively.The P value was calculated using the Kruskal-Wallis test. (B) Distribution of virulence genes in 123 bovine isolates and 185 human isolates. Each number indicatesthe number of different virulence genes found within each group. (C) Frequency of gene clusters among 123 bovine isolates and 185 human isolates. The Pvalues were calculated using the chi-square test. (D) Distribution of virulence genes in newly sequenced clinical K. pneumoniae KpI isolates. Each numberindicates the number of different virulence genes found within each group. (E) Distribution of virulence genes in 123 bovine isolates. Each number indicatesthe number of different virulence genes found within each group.

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 8

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 9: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

bovine isolates. Iron, such as Fe3�, is sequestered by transferrin and lactoferrin (28). Theconcentration of iron in vivo is critical for the control of cellular metabolism in bacteria,and insufficient iron abolishes bacterial growth (29). To grow successfully in hosttissues, bacteria must be able to obtain metal ions from host transport proteins.

The pathogenicity of clinical K. pneumoniae strains is caused by various virulencedeterminants. We found that human isolates carried more virulence genes than bovineisolates. Virulence genes related to siderophore biosynthesis, fimbria proteins, adhe-sion, and secretion systems were found to be prevalent in both bovine and humanisolates. The most common siderophore in cows is enterobactin. However, somesiderophore systems (e.g., aerobactin, colibactin, allantoinase, yersiniabactin, microcin,and salmochelin systems) were significantly more prevalent in human isolates. Further-more, previous studies showed that the ferric uptake operon kfuABC genes weresignificantly associated with tissue invasion in humans (18, 30). Our study confirmedthat kfuABC is also associated with clinical mastitis cases in dairy cows. The secretedsiderophores could grab iron from transferrin, lactoferrin, or ferritin and support thegrowth and infection of K. pneumoniae in hosts (31). Our results provide informationrelated to the genomic structure, the virulence profile, and the resistome of K. pneu-moniae isolates from cows and humans.

MATERIALS AND METHODSSampling design and bacterial strains. This work was conducted between January and August

2017 at four commercial dairy farms, which were eligible for inclusion in the study if they met thefollowing criteria: they had a herd size of �500 lactating cows, they participated in monthly Dairy HerdImprovement Association (DHIA) testing, they had an accurate recording of clinical mastitis cases, andthey were willing to participate. All the farms were located in Cayuga County of New York State, and thefarms were located inside a 15-mi-diameter circular region. Each dairy farm produced its own supply offorages, corn, grass, and alfalfa silage and purchased concentrated feed (corn, soybean meal, etc.) froma local vendor. Additionally, veterinary care was provided by a different veterinary clinic for each one ofthe four farms. Ceftiofur is routinely used on all four dairy farms. A known presence of Klebsiella clinicalmastitis was also required to ensure that an appropriate number of Klebsiella isolates could be collectedfrom each farm. Medians for this sample of herds were 1,057 lactating cows (range, 524 to 1,466 lactatingcows), 12,150 kg of milk/cow per year (range, 11,773 to 13,091 kg of milk/cow per year), and a bulk tanksomatic cell count of 140,500 cells/ml (range, 100,000 to 267,000 cells/ml).

Clinical scores (mild, severe, and culled/dead) for mastitis were assessed after diagnosis according tothe severity scoring system previously described (32). Milk samples were collected from one affectedquarter of mastitis cases aseptically on the farm and cultured using an on-farm culture system (AccuMast;FERA Animal Health, Ithaca, NY) in the diagnostic laboratory of the farm. Plates with Klebsiella-positivecolonies were sent to the laboratory for further analysis.

Antimicrobial susceptibility testing. All Klebsiella isolates were examined for susceptibility toantimicrobial agents using the Bauer-Kirby disk diffusion method following the Clinical and LaboratoryStandards Institute (CLSI) recommendations (33). A double-disk synergy test for screening ESBL-producing isolates was carried out as described previously (34). An additional determination of theceftiofur MIC was carried out for those ESBL-producing isolates. E. coli ATCC 25922 was used as thecontrol strain.

PFGE. Pulsed-field gel electrophoresis (PFGE) was conducted in accordance with the Pulse Netprotocol (www.cdc.gov/pulsenet) provided by the Centers for Disease Control and Prevention (CDC). ThePFGE patterns were analyzed and clustered using BioNumerics (version 7.6) software (Applied Maths,Kortrijk, Belgium). Different PFGE clusters were represented in alphabetical order. Every band differencewithin a PFGE cluster resulted in adding a numerical order to the pulsed-field cluster.

Capsule locus typing. Isolation of genomic DNA from every strain was performed using a PowerSoilDNA isolation kit (MO BIO Laboratories, Inc, Carlsbad, CA) according to the manufacturer’s instructions.A PCR targeting the capsule polysaccharide synthesis (cps) region in K. pneumoniae was used to identifythe capsular types in this study (35). For the strains that had whole-genome sequencing data, we alsoused the Kaptive tool to identify the whole-capsule synthesis locus (K-locus), based on assembly scaffolds(36).

Whole-genome sequencing. In total, 96 K. pneumoniae isolates with distinguishable PFGE profileswere subjected to whole-genome sequencing. DNA libraries were sequenced on a MiSeq 2000 platform(Illumina, Inc., San Diego, CA). The quality of the original reads was evaluated using the FASTQC tool. TheFASTX-trimmer and Trimmomatic tools were used to do the trimming of sequencing reads (37). Thesequencing reads were assembled into contigs using the SPAdes (version 3.10) algorithm (38), andplasmid sequences were assembled using the plasmidSPAdes algorithm (39). Moreover, the genomes of185 K. pneumoniae KpI isolates from humans hospitalized with septicemia, pneumonia, and woundinfection from several countries and 49 K. pneumoniae KpI isolates from bovines were accessed from aprevious study (18). The bovine isolates from the study of Holt et al. (18) included 20 isolates from cowswith clinical mastitis, 10 isolates from cows with subclinical mastitis, and 19 fecal or rumen isolates.

Klebsiella pneumoniae from Cows and Humans Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 9

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 10: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

Phylogenetic analysis. The core genome SNPs were identified by mapping sequence reads againsta reference genome (K. pneumoniae strain NTUH-K2044, GenBank accession no. NC_012731.1) using theBurrows-Wheeler aligner (40). The plasmid sequences in K. pneumoniae strain NTUH-K2044 were ex-cluded. The Genome Analysis Toolkit (GATK) was used for base quality score recalibration, indelrealignment, and duplicate removal (41) and also for SNP and indel discovery, along with genotyping,using standard hard filtering parameters or variant quality score recalibration according to GATK bestpractices recommendations (42, 43). The alleles at all filtered SNP sites were concatenated to form amultiple-sequence alignment of all 308 K. pneumoniae KpI genomes. To construct a maximum likelihoodphylogeny of the K. pneumoniae isolates, the RAxML program was used to calculate the phylogenetic treefrom the combined alignment file with 1,000 bootstraps (44). The lineages of the phylogenetic tree weredefined using the BAPS (version 6.0) program (45). We used Interactive Tree of Life (iTOL) (46) and FigTree(http://tree.bio.ed.ac.uk/software/figtree/) software to visualize and edit the phylogenetic tree.

Genome annotation and taxonomy. For each de novo assembled genome, coding sequences werepredicted using the Prodigal (version 2.6) program (47) and annotated using the rapid prokaryoticgenome annotation tool Prokka (48). We concatenated all ORFs in all genomes into a single file. We thenclustered all the ORFs into clusters that represented a group of highly similar proteins by using theUCLUST algorithm in the USEARCH program with a 90% identity cutoff (49). We extracted the seedsequence for every cluster group in order to do functional annotation, using the DIAMOND program (50),against the Kyoto Encyclopedia of Genes and Genomes (KEGG) database. The descriptive statistics offunctional genes in bovine and humans were calculated in JMP Pro (version 12) software (SAS InstituteInc., Cary, NC) using the chi-square test.

Identification of virulence and AMR genes. We detected known virulence genes and sequenceconservation using a direct read mapping approach with the SRST2 program with a coverage cutoff of90% (51). The reference sequence of virulence genes and alleles was retrieved from the K. pneumoniaeBIGSdb database (https://bigsdb.pasteur.fr/), the virulence factor database (VFDB) (52), and the Patho-genFinder database (53). Taxonomic investigation of virulence among genomic sequences was per-formed with the MG-RAST (version 4.0.2) metagenomics analysis server (http://metagenomics.anl.gov) byuploading the raw reads of sequencing to the server. The associations of virulence genes with clinicalcases were calculated in SAS (version 9.3) software (SAS Institute Inc., Cary, NC).

The AMR genes were identified by use of the SRST2 program (51) according to the ResFinderdatabase (54). The MLST of newly sequenced K. pneumoniae isolates was identified using the MLST(version 1.8) program (55). The new alleles were submitted to the MLST database (https://bigsdb.pasteur.fr/klebsiella/klebsiella.html) to assign new STs. Plasmid replicon types were detected using the Plasmid-Finder (version 1.3) program (56).

Accession number(s). The genome assemblies of newly sequenced strains and the plasmid se-quence of pC5 (GenBank accession no. MF953243) have been deposited in the NCBI database underBioProject no. PRJNA414542 and PRJNA400493, respectively.

SUPPLEMENTAL MATERIALSupplemental material for this article may be found at https://doi.org/10.1128/AEM

.02654-18.SUPPLEMENTAL FILE 1, PDF file, 1 MB.SUPPLEMENTAL FILE 2, XLSX file, 0.3 MB.SUPPLEMENTAL FILE 3, XLSX file, 0.02 MB.SUPPLEMENTAL FILE 4, XLSX file, 0.02 MB.SUPPLEMENTAL FILE 5, XLSX file, 0.05 MB.

ACKNOWLEDGMENTSThis work was supported by Agriculture and Food Research Initiative competitive

grant no. 2013-67015-21233 from the USDA National Institute of Food and Agriculture.We thank the team of curators at the Institut Pasteur MLST system (Paris, France) for

importing novel alleles, profiles, and/or isolates at https://bigsdb.pasteur.fr/, SvetlanaFerreira Lima for help in whole-genome sequencing, Helen Korzec and Martin Zinicolafor collecting the samples, Erika Korzune Ganda for help in data analysis, Qi Sun andMinghui Wang for assistance with genomic analysis, Julie Siler for help in antimicrobialsusceptibility testing, Martin Wiedmann for providing the Salmonella sp. strain H9812used in PFGE typing, Eva Heinz for help in capsule locus identification, and DarylHenderson for help in preparing the manuscript.

REFERENCES1. Magill SS, Edwards JR, Bamberg W, Beldavs ZG, Dumyati G, Kainer MA,

Lynfield R, Maloney M, McAllister-Hollod L, Nadle J, Ray SM, ThompsonDL, Wilson LE, Fridkin SK, Emerging Infections Program Healthcare-Associated Infections and Antimicrobial Use Prevalence Survey Team.

2014. Multistate point-prevalence survey of health care-associatedinfections. N Engl J Med 370:1198 –1208. https://doi.org/10.1056/NEJMoa1306801.

2. Lee W-H, Choi H-I, Hong S-W, Kim K-S, Gho YS, Jeon SG. 2015. Vaccina-

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 10

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 11: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

tion with Klebsiella pneumoniae-derived extracellular vesicles protectsagainst bacteria-induced lethality via both humoral and cellular immu-nity. Exp Mol Med 47:e183. https://doi.org/10.1038/emm.2015.59.

3. Borer A, Saidel-Odes L, Riesenberg K, Eskira S, Peled N, Nativ R, SchlaefferF, Sherf M. 2009. Attributable mortality rate for carbapenem-resistantKlebsiella pneumoniae bacteremia. Infect Control Hosp Epidemiol 30:972–976. https://doi.org/10.1086/605922.

4. Ipekci T, Seyman D, Berk H, Celik O. 2014. Clinical and bacteriologicalefficacy of amikacin in the treatment of lower urinary tract infectioncaused by extended-spectrum beta-lactamase-producing Escherichia colior Klebsiella pneumoniae. J Infect Chemother 20:762–767. https://doi.org/10.1016/j.jiac.2014.08.007.

5. Medrano-Galarza C, Gibbons J, Wagner S, de Passillé AM, Rushen J. 2012.Behavioral changes in dairy cows with mastitis. J Dairy Sci 95:6994 –7002.https://doi.org/10.3168/jds.2011-5247.

6. Bar D, Tauer LW, Bennett G, Gonzalez RN, Hertl JA, Schukken YH, SchulteHF, Welcome FL, Grohn YT. 2008. The cost of generic clinical mastitis indairy cows as estimated by using dynamic programming. J Dairy Sci91:2205–2214. https://doi.org/10.3168/jds.2007-0573.

7. Hertl JA, Grohn YT, Leach JD, Bar D, Bennett GJ, Gonzalez RN, Rauch BJ,Welcome FL, Tauer LW, Schukken YH. 2010. Effects of clinical mastitiscaused by gram-positive and gram-negative bacteria and other organ-isms on the probability of conception in New York State Holstein dairycows. J Dairy Sci 93:1551–1560. https://doi.org/10.3168/jds.2009-2599.

8. Botrel MA, Haenni M, Morignat E, Sulpice P, Madec JY, Calavas D. 2010.Distribution and antimicrobial resistance of clinical and subclinical mas-titis pathogens in dairy cows in Rhone-Alpes, France. Foodborne PathogDis 7:479 – 487. https://doi.org/10.1089/fpd.2009.0425.

9. Erskine RJ, Eberhart RJ, Hutchinson LJ, Spencer SB, Campbell MA. 1988.Incidence and types of clinical mastitis in dairy herds with high and lowsomatic cell counts. J Am Vet Med Assoc 192:761–765.

10. Olde Riekerink RG, Barkema HW, Kelton DF, Scholl DT. 2008. Inci-dence rate of clinical mastitis on Canadian dairy farms. J Dairy Sci91:1366 –1377. https://doi.org/10.3168/jds.2007-0757.

11. Cha E, Bar D, Hertl JA, Tauer LW, Bennett G, Gonzalez RN, Schukken YH,Welcome FL, Grohn YT. 2011. The cost and management of differenttypes of clinical mastitis in dairy cows estimated by dynamic program-ming. J Dairy Sci 94:4476 – 4487. https://doi.org/10.3168/jds.2010-4123.

12. Bannerman DD. 2009. Pathogen-dependent induction of cytokines andother soluble inflammatory mediators during intramammary infection ofdairy cows. J Anim Sci 87:10 –25. https://doi.org/10.2527/jas.2008-1187.

13. Rajala-Schultz PJ, Gröhn YT. 2001. Comparison of economically opti-mized culling recommendations and actual culling decisions of FinnishAyrshire cows. Prev Vet Med 49:29 –39.

14. Todhunter DA, Smith KL, Hogan JS, Schoenberger PS. 1991. Gram-negative bacterial infections of the mammary gland in cows. Am J VetRes 52:184 –188.

15. Erskine RJ, Bartlett PC, VanLente JL, Phipps CR. 2002. Efficacy of systemicceftiofur as a therapy for severe clinical mastitis in dairy cattle. J Dairy Sci85:2571–2575. https://doi.org/10.3168/jds.S0022-0302(02)74340-3.

16. Roberson JR, Warnick LD, Moore G. 2004. Mild to moderate clinicalmastitis: efficacy of intramammary amoxicillin, frequent milk-out, a com-bined intramammary amoxicillin, and frequent milk-out treatment ver-sus no treatment. J Dairy Sci 87:583–592. https://doi.org/10.3168/jds.S0022-0302(04)73200-2.

17. Grohn YT, Gonzalez RN, Wilson D, Hertl JA, Bennett G, Schulte H,Schukken YH. 2005. Effect of pathogen-specific clinical mastitis onherd life in two New York State dairy herds. Prev Vet Med 71:105–125.https://doi.org/10.1016/j.prevetmed.2005.06.002.

18. Holt KE, Wertheim H, Zadoks RN, Baker S, Whitehouse CA, Dance D,Jenney A, Connor TR, Hsu LY, Severin J, Brisse S, Cao H, Wilksch J, GorrieC, Schultz MB, Edwards DJ, Nguyen KV, Nguyen TV, Dao TT, Mensink M,Minh VL, Nhu NT, Schultsz C, Kuntaman K, Newton PN, Moore CE,Strugnell RA, Thomson NR. 2015. Genomic analysis of diversity, popula-tion structure, virulence, and antimicrobial resistance in Klebsiella pneu-moniae, an urgent threat to public health. Proc Natl Acad Sci U S A112:E3574. https://doi.org/10.1073/pnas.1501049112.

19. Sundin GW, Bender CL. 1996. Dissemination of the strA-strBstreptomycin-resistance genes among commensal and pathogenicbacteria from humans, animals, and plants. Mol Ecol 5:133–143.

20. Wang J, Stephan R, Power K, Yan Q, Hachler H, Fanning S. 2014. Nucleotidesequences of 16 transmissible plasmids identified in nine multidrug-resistant Escherichia coli isolates expressing an ESBL phenotype isolated

from food-producing animals and healthy humans. J Antimicrob Che-mother 69:2658 –2668. https://doi.org/10.1093/jac/dku206.

21. Cullik A, Pfeifer Y, Prager R, von Baum H, Witte W. 2010. A novel IS26structure surrounds blaCTX-M genes in different plasmids from Germanclinical Escherichia coli isolates. J Med Microbiol 59:580 –587. https://doi.org/10.1099/jmm.0.016188-0.

22. Dolejska M, Villa L, Hasman H, Hansen L, Carattoli A. 2013. Characteriza-tion of IncN plasmids carrying blaCTX-M-1 and qnr genes in Escherichia coliand Salmonella from animals, the environment and humans. J Antimi-crob Chemother 68:333–339. https://doi.org/10.1093/jac/dks387.

23. Schink AK, Kadlec K, Schwarz S. 2012. Detection of qnr genes amongEscherichia coli isolates of animal origin and complete sequence of theconjugative qnrB19-carrying plasmid pQNR2078. J Antimicrob Che-mother 67:1099 –1102. https://doi.org/10.1093/jac/dks024.

24. Paterson ES, More MI, Pillay G, Cellini C, Woodgate R, Walker GC, Iyer VN,Winans SC. 1999. Genetic analysis of the mobilization and leadingregions of the IncN plasmids pKM101 and pCU1. J Bacteriol 181:2572–2583.

25. Moland ES, Black JA, Hossain A, Hanson ND, Thomson KS, PottumarthyS. 2003. Discovery of CTX-M-like extended-spectrum beta-lactamasesin Escherichia coli isolates from five U.S. states. Antimicrob AgentsChemother 47:2382–2383. https://doi.org/10.1128/AAC.47.7.2382-2383.2003.

26. Lewis JS, II, Herrera M, Wickes B, Patterson JE, Jorgensen JH. 2007. Firstreport of the emergence of CTX-M-type extended-spectrum beta-lactamases (ESBLs) as the predominant ESBL isolated in a U.S. health caresystem. Antimicrob Agents Chemother 51:4015– 4021. https://doi.org/10.1128/AAC.00576-07.

27. Wang G, Huang T, Surendraiah PK, Wang K, Komal R, Zhuge J, Chern CR,Kryszuk AA, King C, Wormser GP. 2013. CTX-M beta-lactamase-producingKlebsiella pneumoniae in suburban New York City, New York, USA. EmergInfect Dis 19:1803–1810. https://doi.org/10.3201/eid1911.121470.

28. Wandersman C, Delepelaire P. 2004. Bacterial iron sources: from sidero-phores to hemophores. Annu Rev Microbiol 58:611– 647. https://doi.org/10.1146/annurev.micro.58.030603.123811.

29. Huang SH, Wang CK, Peng HL, Wu CC, Chen YT, Hong YM, Lin CT. 2012.Role of the small RNA RyhB in the Fur regulon in mediating the capsularpolysaccharide biosynthesis and iron acquisition systems in Klebsiellapneumoniae. BMC Microbiol 12:148. https://doi.org/10.1186/1471-2180-12-148.

30. Ma LC, Fang CT, Lee CZ, Shun CT, Wang JT. 2005. Genomic heterogeneityin Klebsiella pneumoniae strains is associated with primary pyogenic liverabscess and metastatic infection. J Infect Dis 192:117–128. https://doi.org/10.1086/430619.

31. Skaar EP. 2010. The battle for iron between bacterial pathogens andtheir vertebrate hosts. PLoS Pathog 6:e1000949. https://doi.org/10.1371/journal.ppat.1000949.

32. Wenz JR, Garry FB, Barrington GM. 2006. Comparison of disease severityscoring systems for dairy cattle with acute coliform mastitis. J Am VetMed Assoc 229:259 –262. https://doi.org/10.2460/javma.229.2.259.

33. Clinical and Laboratory Standards Institute. 2015. Performance standardsfor antimicrobial susceptibility testing; 25th informational supplement.M100-S25. Clinical and Laboratory Standards Institute, Wayne, PA.

34. Luzzaro F, Mantengoli E, Perilli M, Lombardi G, Orlandi V, Orsatti A,Amicosante G, Rossolini GM, Toniolo A. 2001. Dynamics of a nosocomialoutbreak of multidrug-resistant Pseudomonas aeruginosa producingthe PER-1 extended-spectrum beta-lactamase. J Clin Microbiol 39:1865–1870. https://doi.org/10.1128/JCM.39.5.1865-1870.2001.

35. Pan YJ, Lin TL, Chen YH, Hsu CR, Hsieh PF, Wu MC, Wang JT. 2013.Capsular types of Klebsiella pneumoniae revisited by wzc sequencing.PLoS One 8:e80670. https://doi.org/10.1371/journal.pone.0080670.

36. Wyres KL, Wick RR, Gorrie C, Jenney A, Follador R, Thomson NR, Holt KE.2016. Identification of Klebsiella capsule synthesis loci from whole ge-nome data. Microb Genom 2:e000102. https://doi.org/10.1099/mgen.0.000102.

37. Bolger AM, Lohse M, Usadel B. 2014. Trimmomatic: a flexible trimmer forIllumina sequence data. Bioinformatics 30:2114 –2120. https://doi.org/10.1093/bioinformatics/btu170.

38. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS,Lesin VM, Nikolenko SI, Pham S, Prjibelski AD, Pyshkin AV, Sirotkin AV,Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA. 2012. SPAdes: a newgenome assembly algorithm and its applications to single-cell sequenc-ing. J Comput Biol 19:455– 477. https://doi.org/10.1089/cmb.2012.0021.

39. Antipov D, Hartwick N, Shen M, Raiko M, Lapidus A, Pevzner PA. 2016.

Klebsiella pneumoniae from Cows and Humans Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 11

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 12: Genomic Diversity, Virulence, and Antimicrobial Resistance of ... · ABSTRACT Klebsiella pneumoniae is a leading cause of severe infections in humans and dairy cows, and these infections

plasmidSPAdes: assembling plasmids from whole genome sequencingdata. Bioinformatics 32:3380–3387. https://doi.org/10.1093/bioinformatics/btw493.

40. Li H, Durbin R. 2009. Fast and accurate short read alignment withBurrows-Wheeler transform. Bioinformatics 25:1754 –1760. https://doi.org/10.1093/bioinformatics/btp324.

41. McKenna A, Hanna M, Banks E, Sivachenko A, Cibulskis K, Kernytsky A,Garimella K, Altshuler D, Gabriel S, Daly M, DePristo MA. 2010. TheGenome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res 20:1297–1303. https://doi.org/10.1101/gr.107524.110.

42. DePristo MA, Banks E, Poplin R, Garimella KV, Maguire JR, Hartl C,Philippakis AA, del Angel G, Rivas MA, Hanna M, McKenna A, Fennell TJ,Kernytsky AM, Sivachenko AY, Cibulskis K, Gabriel SB, Altshuler D, DalyMJ. 2011. A framework for variation discovery and genotyping usingnext-generation DNA sequencing data. Nat Genet 43:491– 498. https://doi.org/10.1038/ng.806.

43. Van der Auwera GA, Carneiro MO, Hartl C, Poplin R, Del Angel G,Levy-Moonshine A, Jordan T, Shakir K, Roazen D, Thibault J, Banks E,Garimella KV, Altshuler D, Gabriel S, DePristo MA. 2013. From FastQ datato high confidence variant calls: the Genome Analysis Toolkit bestpractices pipeline. Curr Protoc Bioinformatics 43:11.10.1–11.10.33.https://doi.org/10.1002/0471250953.bi1110s43.

44. Stamatakis A. 2006. RAxML-VI-HPC: maximum likelihood-based phylo-genetic analyses with thousands of taxa and mixed models. Bioinfor-matics 22:2688 –2690. https://doi.org/10.1093/bioinformatics/btl446.

45. Cheng L, Connor TR, Siren J, Aanensen DM, Corander J. 2013. Hierarchi-cal and spatially explicit clustering of DNA sequences with BAPS soft-ware. Mol Biol Evol 30:1224 –1228. https://doi.org/10.1093/molbev/mst028.

46. Letunic I, Bork P. 2016. Interactive Tree of Life (iTOL) v3: an online toolfor the display and annotation of phylogenetic and other trees. NucleicAcids Res 44:W242–W245. https://doi.org/10.1093/nar/gkw290.

47. Hyatt D, Chen GL, Locascio PF, Land ML, Larimer FW, Hauser LJ. 2010.Prodigal: prokaryotic gene recognition and translation initiation site

identification. BMC Bioinformatics 11:119. https://doi.org/10.1186/1471-2105-11-119.

48. Seemann T. 2014. Prokka: rapid prokaryotic genome annotation. Bioin-formatics 30:2068 –2069. https://doi.org/10.1093/bioinformatics/btu153.

49. Edgar RC. 2010. Search and clustering orders of magnitude fasterthan BLAST. Bioinformatics 26:2460 –2461. https://doi.org/10.1093/bioinformatics/btq461.

50. Buchfink B, Xie C, Huson DH. 2015. Fast and sensitive protein alignmentusing DIAMOND. Nat Methods 12:59 – 60. https://doi.org/10.1038/nmeth.3176.

51. Inouye M, Dashnow H, Raven LA, Schultz MB, Pope BJ, Tomita T, ZobelJ, Holt KE. 2014. SRST2: rapid genomic surveillance for public health andhospital microbiology labs. Genome Med 6:90. https://doi.org/10.1186/s13073-014-0090-6.

52. Chen L, Zheng D, Liu B, Yang J, Jin Q. 2016. VFDB 2016: hierarchical andrefined dataset for big data analysis—10 years on. Nucleic Acids Res44:D694 –D697. https://doi.org/10.1093/nar/gkv1239.

53. Cosentino S, Voldby Larsen M, Moller Aarestrup F, Lund O. 2013. Patho-genFinder— distinguishing friend from foe using bacterial whole ge-nome sequence data. PLoS One 8:e77302. https://doi.org/10.1371/journal.pone.0077302.

54. Zankari E, Hasman H, Cosentino S, Vestergaard M, Rasmussen S, Lund O,Aarestrup FM, Larsen MV. 2012. Identification of acquired antimicrobialresistance genes. J Antimicrob Chemother 67:2640 –2644. https://doi.org/10.1093/jac/dks261.

55. Larsen MV, Cosentino S, Rasmussen S, Friis C, Hasman H, Marvig RL,Jelsbak L, Sicheritz-Ponten T, Ussery DW, Aarestrup FM, Lund O. 2012.Multilocus sequence typing of total-genome-sequenced bacteria. J ClinMicrobiol 50:1355–1361. https://doi.org/10.1128/JCM.06094-11.

56. Carattoli A, Zankari E, Garcia-Fernandez A, Larsen MV, Lund O, Villa L,Aarestrup FM, Hasman H. 2014. In silico detection and typing of plas-mids using PlasmidFinder and plasmid multilocus sequence typing.Antimicrob Agents Chemother 58:3895–3903. https://doi.org/10.1128/AAC.02412-14.

Yang et al. Applied and Environmental Microbiology

March 2019 Volume 85 Issue 6 e02654-18 aem.asm.org 12

on February 20, 2021 by guest

http://aem.asm

.org/D

ownloaded from