international journal for parasitology...genetic evidence for the spread of a benzimidazole...

8
Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode Haemonchus contortus Umer Chaudhry a , Elizabeth M. Redman a , Muthusamy Raman b , John S. Gilleard a,a Department of Comparative Biology and Experimental Medicine, Faculty of Veterinary Medicine, University of Calgary, Alberta, Canada b Translational Research Platform for Veterinary Biologicals, Tamil Nadu Veterinary and Animal Sciences University, Chennai 51, India article info Article history: Received 6 March 2015 Received in revised form 28 April 2015 Accepted 29 April 2015 Available online 19 June 2015 Keywords: Haemonchus contortus Isotype-1 -tubulin Anthelmintic drug resistance Benzimidazoles resistance Sheep abstract It is important to understand how anthelmintic drug resistance mutations arise and spread in order to determine appropriate mitigation strategies. We hypothesised that a molecular genetic study of Haemonchus contortus in southern India, a region where resistance may be less advanced than in western Europe and North America, might provide some important insights into the origin and spread of anthel- mintic resistance. The F200Y (TA C) isotype-1 b-tubulin benzimidazole resistance mutation is common in H. contortus throughout the world and the F167Y (TA C) and E198A (GC A) mutations, although less com- mon, have been reported in a number of different countries. We have investigated the haplotypic diver- sity and phylogenetic relationship of isotype-1 b-tubulin benzimidazole resistance alleles for 23 H. contortus populations from small ruminants across southern India. The F200Y (TA C) mutation was most common, being detected in 18/23 populations at frequencies between 9% and 84% and the E198A (GC A) mutation was also detected in 8/23 populations at frequencies between 8% and 18%. The F167Y (TA C) mutation was not detected in any of the 23 populations. Phylogenetic haplotype network analysis sug- gested that the F200Y (TA C) mutation has arisen multiple independent times in the region with at least three independent origins of resistance alleles across the populations surveyed. In contrast, the E198A (GC A) mutation was present on a single haplotype which, given the high level of haplotypic diversity of the susceptible alleles in the region, suggests this particular mutation has spread from a single origin, likely by anthropogenic animal movement. Population genetic analysis of 12 of the H. contortus popula- tions, using a panel of eight microsatellite markers, revealed extremely low genetic differentiation between populations, consistent with the hypothesis of high gene flow among sites. Additionally, there was no significant genetic differentiation between H. contortus taken from sheep and goats which is consistent with H. contortus populations being freely shared between these two different hosts. Overall, we believe these results provide the first clear genetic evidence for the spread of an anthelmintic resistance mutation to multiple different locations from a single origin. Ó 2015 Published by Elsevier Ltd. on behalf of Australian Society for Parasitology Inc. 1. Introduction Anthelmintic resistance in parasitic nematodes is a threat to sustainable livestock production worldwide and an understanding of how resistance mutations arise and spread is important in the development of effective mitigation strategies. However, we still have a limited understanding of these processes at the population level, even for the parasites in which anthelmintic resistance has been most intensively studied such as the gastro-intestinal parasites of small ruminants. Benzimidazole resistance in Haemonchus contortus is an excellent system in which to study the population genetics of anthelmintic resistance due to the avail- able molecular tools and our increasing knowledge of its genetics and population biology (Silvestre and Humbert, 2002; Redman et al., 2008, 2015; Laing et al., 2013). Three resistance-associated non-synonymous single nucleotide polymorphisms (SNPs) in the isotype-1 b-tubulin gene have been associated with benzimidazole resistance in H. contortus. A SNP in codon F200Y (TTC to TA C), resulting in a phenylalanine to tyro- sine substitution, is widespread and often at high frequency in http://dx.doi.org/10.1016/j.ijpara.2015.04.007 0020-7519/Ó 2015 Published by Elsevier Ltd. on behalf of Australian Society for Parasitology Inc. Corresponding author at: Department of Comparative Biology and Experimen- tal Medicine, Faculty of Veterinary Medicine, 3330, Hospital Drive, University of Calgary, Calgary, Alberta T2N 4N1, Canada. Tel.: +1 403 210 6327; fax: +1 403 210 7882. E-mail address: [email protected] (J.S. Gilleard). International Journal for Parasitology 45 (2015) 721–728 Contents lists available at ScienceDirect International Journal for Parasitology journal homepage: www.elsevier.com/locate/ijpara

Upload: others

Post on 10-Mar-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

International Journal for Parasitology 45 (2015) 721–728

Contents lists available at ScienceDirect

International Journal for Parasitology

journal homepage: www.elsevier .com/locate / i jpara

Genetic evidence for the spread of a benzimidazole resistance mutationacross southern India from a single origin in the parasitic nematodeHaemonchus contortus

http://dx.doi.org/10.1016/j.ijpara.2015.04.0070020-7519/� 2015 Published by Elsevier Ltd. on behalf of Australian Society for Parasitology Inc.

⇑ Corresponding author at: Department of Comparative Biology and Experimen-tal Medicine, Faculty of Veterinary Medicine, 3330, Hospital Drive, University ofCalgary, Calgary, Alberta T2N 4N1, Canada. Tel.: +1 403 210 6327; fax: +1 403 2107882.

E-mail address: [email protected] (J.S. Gilleard).

Umer Chaudhry a, Elizabeth M. Redman a, Muthusamy Raman b, John S. Gilleard a,⇑a Department of Comparative Biology and Experimental Medicine, Faculty of Veterinary Medicine, University of Calgary, Alberta, Canadab Translational Research Platform for Veterinary Biologicals, Tamil Nadu Veterinary and Animal Sciences University, Chennai 51, India

a r t i c l e i n f o a b s t r a c t

Article history:Received 6 March 2015Received in revised form 28 April 2015Accepted 29 April 2015Available online 19 June 2015

Keywords:Haemonchus contortusIsotype-1 -tubulinAnthelmintic drug resistanceBenzimidazoles resistanceSheep

It is important to understand how anthelmintic drug resistance mutations arise and spread in order todetermine appropriate mitigation strategies. We hypothesised that a molecular genetic study ofHaemonchus contortus in southern India, a region where resistance may be less advanced than in westernEurope and North America, might provide some important insights into the origin and spread of anthel-mintic resistance. The F200Y (TAC) isotype-1 b-tubulin benzimidazole resistance mutation is common inH. contortus throughout the world and the F167Y (TAC) and E198A (GCA) mutations, although less com-mon, have been reported in a number of different countries. We have investigated the haplotypic diver-sity and phylogenetic relationship of isotype-1 b-tubulin benzimidazole resistance alleles for 23 H.contortus populations from small ruminants across southern India. The F200Y (TAC) mutation was mostcommon, being detected in 18/23 populations at frequencies between 9% and 84% and the E198A (GCA)mutation was also detected in 8/23 populations at frequencies between 8% and 18%. The F167Y (TAC)mutation was not detected in any of the 23 populations. Phylogenetic haplotype network analysis sug-gested that the F200Y (TAC) mutation has arisen multiple independent times in the region with at leastthree independent origins of resistance alleles across the populations surveyed. In contrast, the E198A(GCA) mutation was present on a single haplotype which, given the high level of haplotypic diversityof the susceptible alleles in the region, suggests this particular mutation has spread from a single origin,likely by anthropogenic animal movement. Population genetic analysis of 12 of the H. contortus popula-tions, using a panel of eight microsatellite markers, revealed extremely low genetic differentiationbetween populations, consistent with the hypothesis of high gene flow among sites. Additionally, therewas no significant genetic differentiation between H. contortus taken from sheep and goats which isconsistent with H. contortus populations being freely shared between these two different hosts.Overall, we believe these results provide the first clear genetic evidence for the spread of an anthelminticresistance mutation to multiple different locations from a single origin.

� 2015 Published by Elsevier Ltd. on behalf of Australian Society for Parasitology Inc.

1. Introduction

Anthelmintic resistance in parasitic nematodes is a threat tosustainable livestock production worldwide and an understandingof how resistance mutations arise and spread is important in thedevelopment of effective mitigation strategies. However, we stillhave a limited understanding of these processes at the population

level, even for the parasites in which anthelmintic resistance hasbeen most intensively studied such as the gastro-intestinalparasites of small ruminants. Benzimidazole resistance inHaemonchus contortus is an excellent system in which to studythe population genetics of anthelmintic resistance due to the avail-able molecular tools and our increasing knowledge of its geneticsand population biology (Silvestre and Humbert, 2002; Redmanet al., 2008, 2015; Laing et al., 2013).

Three resistance-associated non-synonymous single nucleotidepolymorphisms (SNPs) in the isotype-1 b-tubulin gene have beenassociated with benzimidazole resistance in H. contortus. A SNP

in codon F200Y (TTC to TAC), resulting in a phenylalanine to tyro-sine substitution, is widespread and often at high frequency in

Page 2: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

722 U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728

many countries (Silvestre and Humbert, 2002; Hoglund et al.,2009; Brasil et al., 2012; Redman et al., 2015). A SNP at codon

F167Y (TTC to TAC), resulting in a phenylalanine to tyrosine substi-tution, has also been detected in a number of countries although itis generally considered to be less common than the F200Y (TTC to

TAC) mutation (Silvestre and Cabaret, 2002; de Lourdes Mottierand Prichard, 2008; Barrère et al., 2012; Brasil et al., 2012;Redman et al., 2015). However, it is noteworthy that the F167Y

(TAC) mutation was present in six out of seven H. contortus popu-lations examined in a recent UK study and its frequency across

these six farms was similar to that of the F200Y (TAC) mutationoverall (Redman et al., 2015). Indeed, in one population where

the F200Y (TAC) mutation was detected at very low frequency,

the F167Y (TAC) mutation was present almost at fixation (>95%

frequency). Hence, the F167Y (TAC) mutation may be of similar

importance as the F200Y (TAC) mutation for benzimidazole resis-tance of H. contortus in the UK. A third SNP at codon E198A (GAA

to GCA) resulting in a glutamate to alanine substitution has beendetected in just two field-derived laboratory passaged populationsof H. contortus to date; one from South Africa and one fromAustralia (Ghisi et al., 2007; Rufener et al., 2009; Kotze et al.,

2012). It appears the E198A (GCA) mutation is the rarest of thethree benzimidazole mutations globally, although it has not beenthe focus of many published studies to date.

In our recent population genetic study of H. contortus on seven

UK farms, we presented evidence that both the F200Y (TAC) and

the F167Y (TAC) mutations present on those farms had multipleindependent origins (Redman et al., 2015). This suggests that ben-zimidazole resistance mutations have repeatedly arisen in the UK,although we still have a poor understanding of how often this hasoccurred and of the relative importance of recurrent versuspre-exisiting mutations (Redman et al., 2015). We are investigatingthis question in a number of different countries and southern Indiawas chosen on the basis that resistance might be at an earlier stagethan in many of the regions studied to date. Anthelmintic treat-ment of livestock is generally less regimented in rural southernIndia than it is in western Europe and North America, and genericdrug formulations of unknown efficacy are often used. In addition,we suggest that the warm, humid climate would result in a largerefugia population and reduce the rate at which resistancedevelops. There are only a few published studies on anthelminticresistance from India but benzimidazole resistance has beendemonstrated by fecal egg count reduction tests and the F200Y

(TAC) mutation has been detected by allele-specific PCR in north-ern India (Tiwari et al., 2006; Chandra et al., 2014). However, theoverall situation is unclear.

In this paper, we present evidence for the spread of the E198A

(GCA) mutation from a single origin to multiple locations in south-

ern India. In contrast, the F200Y (TAC) mutation appears to havemultiple independent origins, consistent with its more frequentoccurrence. The results presented provide, to our knowledge, thefirst known genetic evidence of the spread of an anthelmintic resis-tance mutation from a single source, presumably due to the largeamount of anthropogenic livestock movement in the region.

2. Materials and methods

2.1. Harvesting of adult Haemonchus worms

Adult Haemonchus worms were harvested from the abomasa of23 individual ruminant hosts that were slaughtered at 10 differentabattoirs located across four different provinces of southern Indiain 2011 (Tamil Nadu, Andhra Pradesh, Kerala and Karnataka)

(Fig. 1, Supplementary Table S1). Each animal that was necropsiedin each abattoir was known to originate from a different pas-ture/village in the region, although the precise location was notavailable. Hence, each of the 23 animals harboured geographicallydistinct parasite populations distributed across the four provinces.Twelve of the H. contortus populations were collected from sheep;two from the Chennai abattoir (Pop2S and Pop1S), two from theSalem abattoir (Pop51S and Pop52S), one from the Tirunelveliabattoir (Pop16S), two from the Madurai abattoir (Pop41S andPop42S), two from the Vijayawada abattoir (Pop71S and Pop72S)and three from the Tirupathia abattoir (Pop27S, Pop28S andPop26S). Eight of the H. contortus populations were collected fromgoats; two from the Dharmapuri abattoir (Pop11G and Pop12G),two from the Trichy abattoir (Pop21G and Pop22G), two from theCochin abattoir (Pop32G and Pop31G) and two from theMangalore abattoir (Pop61G and Pop62G). Two of the H. contortuspopulations were collected from cattle, one each from theVijayawada (Pop2C) and Cochin abattoirs (Pop3C) and a singleH. contortus population was collected from a buffalo at theChennai abattoir (Pop41B) (Fig. 1, Supplementary Table S1).

2.2. Genomic DNA isolation

Adult worms were fixed in 70% ethanol immediately followingremoval from the host abomasum. In order to avoid contaminationwith any progeny DNA, the heads of male and female worms weredissected and lysed in single 0.2 ll tubes containing 50 ll of pro-teinase K lysis buffer and stored at �80 �C as previously described(Redman et al., 2008; Chaudhry et al., 2015). One microliter of a 1:5dilution of neat single worm lysate was used as PCR template andidentical dilutions of lysate buffer, made in parallel, were used asnegative controls. To prepare pooled lysates of each population,1 ll aliquots of each individual neat adult worm head lysate werepooled. One microliter of a 1:20 dilution of pooled lysate was usedas template for subsequent PCRs.

2.3. Pyrosequence genotyping of the rDNA ITS-2 P24 SNP to determineHaemonchus spp. identity

A 321 bp fragment encompassing the rDNA ITS-2 region wasPCR amplified from individual H. contortus adult worm lysatesusing a ‘‘universal’’ forward primer complementary to the 5.8SrDNA coding sequence (NC1F: 50-ACG TCT GGT TCA GGG TTGTT-30) and a biotin-labeled reverse primer complementary to the28S rDNA coding sequence (NC2R: 50-Biotin-TTA GTT TCT TTTCCT CCG CT-30) (Stevenson et al., 1995; Chaudhry et al., 2015).The SNP at position 24 of the ITS-2 (which discriminates betweenH. contortus and Haemonchus placei) was determined by a previ-ously described pyrosequence genotyping assay using sequencingprimer 50-CATATACTACAATGTGGCTA-30 (Chaudhry et al., 2015). Aminimum of 18 and a maximum of 45 individual adult worms(850 worms in total) were genotyped per population (23 popula-tions in total) (Supplementary Table S1). All the worms genotypedin the 23 populations described above were confirmed as H. contor-

tus (P24; A genotype) (Chaudhry et al., 2015).

2.4. Pyrosequence genotyping to determine the relative frequenciesof the isotype-1 b-tubulin P200, P198 and P167 benzimidazoleresistance-associated SNPs in H. contortus populations

A 328 bp fragment, spanning exons 4 and 5 and the interveningintron of the isotype-1 b-tubulin gene was PCR amplified frompooled DNA lysates for each of the 23 H. contortus populations(Supplementary Table S1). Final PCR conditions used 1� thermopolreaction buffer, 2 mM MgSO4, 100 lM dNTPs, 0.1 lM forward andreverse primers and 1.25 U Taq DNA polymerase (New England

Page 3: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

Fig. 1. Relative frequencies of the F200Y (TAC), F167Y (TAC) and E198A (GCA) isotype-1 b-tubulin benzimidazole resistance-associated mutations in Haemonchus contortuspopulations from southern India. Each H. contortus population is represented by three pie charts with one for each resistance mutation, F200Y (TTC/TAC), F167Y (TTC/TAC)and E198A (GAA/GCA). Each of these pie charts shows the relative frequency of the resistant versus susceptible single nucleotide polymorphism based on allele quantificationby pyrosequence genotyping of pooled DNA from between 18 and 44 worms per population (Supplementary Table S1). The resistance-associated single nucleotidepolymorphism genotype frequency – F200Y (TAC) or E198A (GCA) – is shown in light gray hatch and ‘‘susceptible’’ SNP genotype frequency – F200Y (TTC), F167Y (TTC) andE198A (GAA) – is shown in black. Geographic locations of abattoirs are indicated with small circles on the map and the abattoir names from which the samples were obtainedare shown above the labeling lines. The four provinces are indicated on the map: (A) Tamil Nadu, (B) Andhra Pradesh, (C) Karnataka, and (D) Kerala.

U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728 723

Biolabs, USA) and a previously published primer pair was used;forward primer (HcPYRF: 50-GAC GCA TTC ACT TGG AGG AG-30)and reverse primer (HcPYRR: 50-Biotin-CAT AGG TTG GATTTGTGA GTT-30) (von Samson-Himmelstjerna et al., 2009). Thethermo-cycling parameters consisted of an initial 95 �C for 5 minfollowed by 35 cycles of 95 �C for 1 min, 53 �C for 1 min and72 �C for 1 min and a single final extension cycle of 72 �C for 5 min.

Following PCR amplification, the relative frequency of the

F167Y (TAC), E198A (GCA) and F200Y (TAC) isotype-1 b-tubulinSNPs in the amplicons derived from each parasite population weredetermined by pyrosequence genotyping using the PryoMark IDsystem using the allele quantification (AQ) mode (Biotage,Sweden). Previously published sequencing primers were used for

the F167Y (TAC) and F200Y (TAC) mutations; Hcsq167: 50-ATAGAA TTA TGG CTT CGT-30 and Hcsq200: 50-TAG AGA ACA CCGATG AAA CAT-30, respectively (Hoglund et al., 2009; vonSamson-Himmelstjerna et al., 2009). A new sequencing primer,Hcsq198: 50-ACT GGT AGA GAA CAC CG-30, was used for the

E198A (GCA) mutation. The base dispensation was set toGATGCTCGT for codon 167, GATGCAGCA for codon 198 andGATGCTGTA for codon 200. Peak heights were measured usingthe AQ mode of the PSQ 96 single nucleotide position software.In order to validate the use of peak heights to estimate relativeSNP frequencies in amplicons derived from pooled DNA, between24 and 41 individual worms from nine of the populations were alsoindividually genotyped at the three SNPs (see Supplementary

Table S2). The frequency of the F200Y (TAC) SNP, based on pooleddata, correlated well with that derived from the individual wormgenotyping for these nine populations (see Supplementary

Fig. S1). The E198A (GCA) SNP was only present at low frequencyin the populations and thus the correlation was less precise.Nevertheless the pooled data still provided a sufficient estimateof the allele frequency appropriate for the work undertaken in thisstudy (see Supplementary Fig. S1).

2.5. Sequencing of cloned H. contortus isotype-1 b-tubulin

The same 328 bp isotype-1 b-tubulin fragment described inSection 2.4 was also amplified, cloned and sequenced from thepooled lysates from 13 of the populations (Pop28S, Pop1S, Pop2S,Pop2S, Pop16S, Pop3C, Pop26S, Pop27S, Pop11G, Pop12G, Pop51S,Pop31G and Pop32G). Primers HcPYRF: 50-GAC GCA TTC ACT TGGAGG AG-30 and HcPYRR: 50-CAT AGG TTG GAT TTGTGA GTT-30 wereagain used but with the following PCR reaction conditions: 5�Phusion HF reaction buffer, 2 mM MgSO4, 200 lM dNTPs, 0.2 lMforward and reverse primers and 1 U of Phusion high fidelity DNApolymerase (Finenzyme Inc., USA). The thermo-cycling parametersof isotype-1 b-tubulin consisted of an initial 98 �C for 30 s followedby 35 cycles of 98 �C for 10 s, 59 �C for 30 s and 72 �C for 1 min witha single final extension cycle of 72 �C for 5 min. Amplicons werecloned into a PJET 1.2/BLUNT vector (Thermo Scientific, USA) andsequenced using previously described procedures (Chaudhryet al., 2014). Sequences were aligned with H. contortus isotype-1b-tubulin sequences (GenBank Accession Number X67489) and edi-ted using Geneious Pro 5.4 software (version 5.6)

2.6. Bioinformatic filtering of H. contortus isotype-1 b-tubulinsequences to remove PCR induced mutations

Polymorphisms appearing more than once in the sequence dataset are expected to be real, whereas polymorphisms that onlyoccur once are possible artefacts due to polymerase-inducederrors. We used a previously described method to test this forour data (Redman et al., 2015). The distribution of the SNPs, whenplotted along the isotype-1 -tubulin gene model, strongly supportsthe hypothesis that the majority of unique SNPs were PCR-inducedmutations whereas those appearing more once were genuine poly-morphisms (Supplementary Fig. S2). Specifically, SNPs appearingmore than once were either clustered within the introns or were

Page 4: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

724 U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728

synonymous changes in the exons (except for those occurring atthe resistance-associated positions, P198 and P200). This is thepattern expected for genuine polymorphisms. In contrast, theSNPs that occurred only once in the entire dataset were evenly dis-tributed across intron and exons with no bias for synonymousmutations in the exons. Consequently, we only considered SNPsoccurring more than once in the dataset in order to take a conser-vative approach and ensure only real polymorphisms were consid-ered. This resulted in 21 different H. contortus haplotypes in thefinal filtered dataset which were used for all following analyses.For completeness, we also constructed trees and haplotype net-works with the full unfiltered dataset and, other than increasingthe total number of haplotypes, this did not change any of the phy-logenetic relationships or overall conclusions (data not shown).

2.7. Phylogenetic network analysis of H. contortus isotype-1 b-tubulinsequence data

Networks based on genetic distance were computed using theneighbor-net method employed in SplitsTrees4 (Huson andBryant, 2006) to produce circular (equal angle) split networks.Median joining networks were generated in Network 4.6.1 soft-ware (Fluxus Technology Ltd). A full median network containingall possible shortest trees was generated by setting the epsilonparameter equal to the greatest weighted distance (epsilon = 10).All unnecessary median vectors and links were removed with theMaximum Parsimony (MP) option (Polzin and Daneschmand,2003). A phylogenetic network tree of the isotype-1 -tubulin hap-lotypes was reconstructed using maximum Likelihood (ML) inMEGA5 (Tamura et al., 2011). The program jModeltest 12.2.0(Posada, 2008) was used to select the appropriate model of nucleo-tide substitutions for ML analysis. According to the Bayesian infor-mation criterion, the best scoring was Hasegawa–Kishino–Yano(HKY+G). The model of substitution was used with parametersestimated from the data. Branch supports were obtained by 1000bootstraps of the data. The most probable ancestral node wasdetermined by rooting the networks to a closely related outgroupand a Teladorsagia circumcincta sequence was used to root the H.contortus network.

2.8. Microsatellite genotyping and population genetic analysis

Previously described H. contortus microsatellite markers (Otsenet al., 2000; Redman et al., 2008, 2015) were screened to produce apanel of eight markers that consistently amplified loci from theIndian populations under study (data not shown). A summary ofprimer sequences and allele ranges are given in SupplementaryTable S3 and the PCR conditions for each marker were as previ-ously described by Redman et al. (2008). The forward primer ofeach microsatellite primer pair was 50 end labeled with fluorescentdye (IDT, Canada) and the GeneScan ROX 400 internal size stan-dard was used on the ABI Prism 3100 genetic analyser (AppliedBiosystems, USA). Individual chromatograms were analysed usingGene Mapper software version 4.0 to accurately size the ampliconsand determine genotypes (Applied Biosystems).

From the multi-locus microsatellite genotype data, heterozy-gosity (He and Ho), allele richness and estimates of FIS for eachlocus were calculated using Arlequin 3.1 (Excoffier et al., 2005).The exact test of Guo and Thompson (1992) was used to statisti-cally evaluate deviations from Hardy–Weinberg equilibrium forall populations (Guo and Thompson, 1992). Significance levelswere adjusted using the sequential method of Bonferroni for mul-tiple comparisons in the same dataset (Rice, 1989). There was asignificant departure from Hardy–Weinberg equilibrium, evenafter Bonferroni correction, in addition to relatively high FIS valuesfor 10 out of the 108 loci combinations for H. contortus

(Supplementary Tables S3). The presence of null alleles formicrosatellite loci has been previously reported and is the likelyreason for these departures from Hardy–Weinberg Equilibrium(Grillo et al., 2007; Hunt et al., 2008; Redman et al., 2008;Silvestre et al., 2009). Linkage disequilibrium analysis was carriedout by GENEPOP version 3.3 (Rousset and Raymond, 1995) usinga log likelihood ratio test statistic (G-test). Partitions of microsatel-lite diversity between and within populations were estimatedthrough an analysis of molecular variance (AMOVA) (Excoffieret al., 1992). Data were defined as ‘standard’ rather than ‘mi-crosatellite’, as loci did not necessarily adhere to the stepwisemutation model. Pairwise FST values were calculated and signifi-cance testing was undertaken by random permutation inArlequin 3.11.

3. Results

3.1. The F200Y and E198A but not the F167Y benzimidazole resistance-associated isotype-1 b-tubulin polymorphisms were commonly foundin H. contortus populations across southern India

A 328 bp fragment of the isotype-1 b-tubulin gene was PCR ampli-fied from 23 H. contortus populations across southern India and therelative frequencies of the three currently known benzimidazole

resistance-associated SNPs, F167Y (TAC), E198A (GCA) and F200Y

(TAC), were determined by pyrosequence genotyping (using poolsof 18–45 worms per population; see Supplementary Table S1).Benzimidazole resistance-associated SNPs were found in 18 of the

23 populations with the F200Y (TAC) and E198A (GCA) SNPs being

identified in multiple populations. The F200Y (TAC) and the E198A

(GCA) SNPs were detected in 18 and eight populations at frequenciesbetween 9–84% and 8–18%, respectively (Fig. 1, SupplementaryTable S1). The benzimidazole resistance-associated SNP F167Y

(TAC) was not detected in any of the populations.

3.2. Sequence diversity and phylogenetic relationships of isotype-1 -tubulin alleles in H. contortus from southern India

The 328 bp isotype-1 b-tubulin fragment was PCR amplified,cloned and sequenced from pooled DNA from each of 13 popula-tions. For each population, template DNA was pooled frombetween 18 and 45 worms (Supplementary Table S1) and theinserts of at least 10 clones were sequenced per population.Following bioinformatic filtering (see Section 2.6), a total of 21 dif-ferent H. contortus isotype-1 -tubulin haplotypes (GenBank acces-sion numbers KP792514–KP792534) were identified in the 159sequences generated from the 13 populations. Seven out of the

21 haplotypes encoded the F200Y (TAC) resistance polymorphism

and one encoded the E198A (GCA) resistance polymorphism. ASplitsTrees network was constructed with all 21 isotype-1 -tubulinhaplotypes to examine their phylogenetic relationships (Fig. 2).

The F200Y (TAC) resistance polymorphism was present on haplo-types located in three distinct parts of the network, each of whichalso contained at least one susceptible haplotype. In contrast, the

E198A (GCA) resistance polymorphism was present on a singlehaplotype in the network. A ML tree was also constructed and showedthe same phylogenetic relationships (Supplementary Fig. S4).

3.3. Geographical distribution of F200Y (TAC) and E198A (GCA)isotype-1 -tubulin benzimidazole resistance haplotypes in southernIndia

Frequency histograms of both the susceptible and resistant hap-lotypes showed that most of the populations contained multiple

Page 5: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

Fig. 2. Split Trees network of the Haemonchus contortus isotype-1 -tubulin sequences generated with the neighbor-net method of SplitsTrees4 (Huson and Bryant, 2006). Thecircles in the network represent the different haplotypes and the size of each circle is proportional to the frequency in the population. The haplotypes containing the differentmutations are shaded as follows: susceptible haplotypes containing F200Y (TTC)/F167Y (TTC)/E198A (GAA) are black; P200Y-resistant haplotypes containing F200Y (TAC)/F167Y (TTC)/E198A (GAA) are light hatch; the P198A-resistant haplotype (HR24) containing F200Y (TTC)/F167Y (TTC)/E198A (GCA) is shown by vertical line shading.

U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728 725

different resistance haplotypes (Supplementary Fig. S3). The onlyexceptions were populations Pop31G, Pop32G and Pop1S in which

the F200Y (TAC) resistance-associated polymorphism was detectedon a single resistance haplotype (Supplementary Fig. S3). TheMedian-Joining network was constructed to integrate the geo-graphical distribution and phylogenetic relationships of the differ-ent isotype-1 -tubulin haplotypes (Fig. 3). The Median-Joiningnetwork was congruent with the SplitsTrees network (Fig. 2) andML tree (Supplementary Fig. S4) analysis. The only differencewas in the relative position of haplotype Hr25 in the network.

The two F200Y (TAC) resistance-associated haplotypes that arepresent at the highest frequencies in the dataset (Hr23 and Hr34)were present in a large number of different parasite populationsacross southern India (11 and seven different populations, respec-

tively). The only F200Y (TAC) resistance-associated haplotypes thatwere present in just a single population were those present at low

frequencies in the data set overall. In contrast, the E198A (GCA)resistance-associated mutation was present on only a single haplo-type in the network (Hr24) for all six of the populations in which itwas sequenced (Fig. 3).

3.4. Genetic diversity and population structure of H. contortus insouthern India

A panel of eight microsatellites was used to genotype 18 indi-vidual worms in each of 12 populations in order to investigate thebasic population genetic structure of H. contortus in southernIndia. Six of the populations (Pop26S, Pop27S, Pop51S, Pop52S,

Pop71S and Pop22S) were from sheep and six from goats(Pop11G, Pop12G, Pop31G, Pop32G, Pop61G and Pop62G). Eight

of the populations contained both the F200Y (TAC) and E198A

(GCA) resistance-associated SNPs (Pop11G, Pop12G, Pop22S,Pop26S, Pop27S, Pop32G, Pop52S and Pop62G), two populations

just the F200Y (TAC) resistance-associated SNP (Pop31G andPop51S) and two populations did not contain anyresistance-associated SNPs (Pop61G and Pop71S). There were nomajor departures from linkage equilibrium for any particularcombination of loci across all populations, indicating that allelesat these loci were randomly associating and not genetically linked(data not shown). All populations were polymorphic at all loci,with the number of alleles per locus ranging from two to 13(Supplementary Table S4). Haemonchus contortus showed a highlevel of overall genetic diversity in all populations; the meanallele richness was 6.565 ± 0.371 alleles per locus with anexpected heterozygosity (He) of 0.6449 (range: 0.651–0.6900).There was very little difference in overall genetic diversitybetween the 12 populations. AMOVA analysis estimated thatthe percentage of variation that partitioned between 12 popula-tions was 0.006%, suggesting extremely low geneticsub-structuring of H. contortus in southern India. This wasreflected by the very low pairwise FST estimates calculatedbetween each of the 12 populations. Only five out of 66 possiblepairwise comparisons showed significant differentiation (Table 1)based on FST values and, even in these cases, the FST values werevery low, ranging between 0.0244 (between Pop31G and Pop52S)and 0.0351 (between Pop61G and Pop22S).

Page 6: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

Fig. 3. Median joining network of the Haemonchus contortus isotype-1 -tubulin sequences generated in Network 4.6.1 (Fluxus Technology Ltd.). A full median networkcontaining all possible shortest trees was generated by setting the epsilon parameter equal to the greatest weighted distance (epsilon = 10). All unnecessary median vectorsand links were removed with the Maximum Parsimony option (Polzin and Daneschmand, 2003). The size of the circle representing each haplotype is proportional to itsfrequency in the dataset and the colors in each circle reflect the frequency distribution in each population as indicated in the color key in the inset map. Small red dotsrepresent median vectors. The number of mutations separating adjacent sequence nodes or median vectors is indicated along connecting branches and the lengths of the linesconnecting the haplotypes are proportional to the number of nucleotide changes. The most probable ancestral node is determined by rooting the network to a closely relatedoutgroup, Teladorsagia circumcincta (GenBank accession number FN599034). The text providing the name of each haplotype is color coded as follows: susceptible haplotypeF200Y (TTC)/F167Y (TTC)/E198A (GAA) is in black text; P200Y-resistant haplotype F200Y (TAC)/F167Y (TTC)/E198A (GAA) is in blue text; P198A-resistant haplotype F200Y(TTC)/ F167Y (TTC)/E198A (GCA) is in pink text.

726 U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728

4. Discussion

The frequency with which anthelmintic drug resistance arisesand the extent to which it spreads are important considerationsfor its prevention and management (Silvestre and Humbert, 2002;Skuce et al., 2010; Redman et al., 2015). Previous work characteriz-ing benzimidazole resistance mutations in H. contortus allows us tostudy the origin and spread of resistance in a much more targetedway than for other drug classes and parasite species. In this paper,we present a molecular genetic study of benzimidazole resistancein H. contortus in southern India. We chose this region for two mainreasons. Firstly, small ruminants in rural areas of southern India aretreated in a sporadic manner, often with generic drugs of unknownquality. Also, the warm, humid climate is expected to allow a

relatively large refugia population of free-living larvae on pasturefor most of the year. These factors are expected to lead to lessintense selection pressure overall and thus led us to hypothesisethat benzimidazole resistance might be at an earlier stage in south-ern India than in many of the countries examined to date. Secondly,although the effect of climate on refugia is complex, the warmhumid climate is expected to allow relatively large refugia popula-tions. We believed that studying benzimidazole resistance in such aregion should provide some interesting insights into the how resis-tance emerges and spreads, and how this can be examined usingmolecular genetic approaches.

Although benzimidazole resistance is already widespread insouthern India, our results suggest it is indeed less advanced thanin those European and North American countries examined to date

Page 7: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

Table 1Pairwise FST values based on genotyping 18 individual worms from each of 12 Haemonchus contortus populations with eight microsatellite markers. Pairwise comparisons withstatistically significant genetic differentiation (P < 0.002) are highlighted in bold.

Pop11G Pop12G Pop26S Pop27S Pop31G Pop32G Pop51S Pop52S Pop61G Pop62G Pop71S

Pop12G �0.0101Pop26S 0.0030 0.0041Pop27S �0.0010 0.0034 0.0132Pop31G �0.0095 �0.006 �0.0035 0.0059Pop32G �0.0020 �0.0086 �0.0057 0.0123 0.0046Pop51S �0.0035 0.0003 �0.0041 0.0201 �0.0099 0.0089Pop52S �0.0040 0.0040 0.0248 0.0268 0.0244 0.0062 0.0211Pop61G �0.0174 0.0021 0.0123 0.0110 0.0019 0.0009 0.0061 0.0012Pop62G �0.0075 0.0059 0.0038 0.0079 �0.0052 0.0031 0.0045 0.0137 0.0072Pop71S �0.0051 �0.0069 �0.0052 0.0253 �0.0070 �0.0002 �0.0109 0.0087 0.0029 0.0056Pop22S 0.0157 0.0053 0.0008 0.0299 0.0055 0.0102 0.0118 0.0347 0.0351 0.0164 0.0002

U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728 727

(Silvestre and Cabaret, 2002; de Lourdes Mottier and Prichard,2008; Barrère et al., 2012, 2013a,b; Brasil et al., 2012; Redmanet al., 2015; Chaudhry, U., 2015. Molecular analysis of benzimida-zole resistance in Haemonchus contortus and Haemonchus placei.PhD Thesis, University of Calgary, Canada). Although the

isotype-1 b-tubulin P200Y (TAC) mutation was present in 18 outof 23 populations, it was still at a relatively low frequency (<25%)

in over half of these (13/23). Similarly, although the E198A (GCA)mutation was present in 8/23 populations, it was at a low frequencyin all cases (maximum frequency of 18%). It is interesting that the

E198A (GCA) mutation was identified in so many populations giventhat it has only been previously reported in just two H. contortusfield-derived isolates (from Australia and South Africa) and wasnot detected in surveys of benzimidazole-resistant H. contortusfrom the UK, Canada and USA (Ghisi et al., 2007; Rufener et al.,2009; Barrère et al., 2012, 2013a,b; Kotze et al., 2012; Chaudhry,

2015, PhD Thesis, cited earlier). Hence the E198A (GCA) mutationappears to be much more common in southern India than in theother countries examined to date, even though benzimidazole

resistance appears less advanced based on the P200Y (TAC) muta-tion distribution and frequency overall (Hoglund et al., 2009;Barrère et al., 2012, 2013a; Redman et al., 2015; Chaudhry, 2015,PhD Thesis, cited earlier). In contrast, the opposite applies to the

F167Y (TAC) mutation. We did not identify this mutation in anyof the 23 southern Indian populations, despite this mutation beingfound in most other countries examined to date (Silvestre andCabaret, 2002; de Lourdes Mottier and Prichard, 2008; Barrèreet al., 2012, 2013a,b; Brasil et al., 2012). Indeed, in our recent UK

study, the F167Y (TAC) mutation was found in H. contortus on sixout of seven farms examined and was at a similar overall frequency

to the P200Y (TAC) mutation. The contrast between southern India

and the UK is striking; the F167Y (TAC) mutation was commonly

found in the UK but the E198A (GCA) mutation was not identified.However, the exact opposite was the case for southern India. Hence,

although the P200Y (TAC) mutation seems to be inevitably presentin H. contortus populations throughout the world, there appears tobe marked regional variation in the relative occurrence of the F167Y

(TAC) and E198A (GCA) mutations. The most parsimonious expla-nation of these observed patterns of occurrence is that the F167Y

(TAC) and E198A (GCA) mutations rarely appear in a region butthen become widespread by migration. In this model, it is unsur-prising that a particular mutation would be widespread in somecountries or regions but absent in others.

In the case of the E198A (GCA) mutation in southern India, thephylogenetic analysis of the resistance haplotypes also supportsthe hypothesis that it rarely arises. Indeed, in this case, it appearsthat this mutation has spread to multiple different locations insouthern India from a single origin. The mutation is present on just

a single haplotype in all six of the H. contortus populations fromwhich it was sequenced (Figs. 2 and 3). In contrast, there was a largeamount of sequence diversity in susceptible haplotypes evenwithin the short fragment that was sequenced (13 different suscep-tible haplotypes in the dataset). Hence, given this high level of sus-ceptible haplotype diversity in southern India, it would be

extremely unlikely that the E198A (GCA) mutation repeatedly aroseonly on the same haplotype (Hr24). Hence, the evidence strongly

suggests that the E198A (GCA) mutation in the populations exam-ined derives from a single origin. We believe that this representsthe first clear genetic evidence of the spread of an anthelminticresistance mutation across multiple locations in a country.

In contrast, the P200Y (TAC) mutation was present on multiple,phylogenetically diverse, haplotypes and their phylogenetic rela-tionship suggests there are at least three independent origins of

the P200Y (TAC) mutation in the populations examined (Figs. 2and 3). Within the haplotype networks, each of three groups ofresistance haplotypes (Hr23/Hr27/Hr36; Hr22/Hr35; Hr25/Hr34)were more closely related to one or more susceptible haplotypesin the network than they were to any of the other resistant haplo-types (Figs. 2 and 3). This provides strong evidence that these areindependently derived alleles. Hence, the phylogenetic evidence

supports the view that the P200Y (TAC) mutation originates much

more commonly in parasite populations than the E198A (GCA)

mutation. This helps to explain why the P200Y (TAC) mutation isseen more commonly and consistently across the world than the

E198A (GCA) mutation whose distribution is much more highlyvariable as discussed above.

We examined the population structure of 12 H. contortus popu-lations from southern India using a panel of eight highly polymor-phic microsatellite markers. The diversity within each populationwas very high as has been previously reported for this parasite(Blouin et al., 1995; Silvestre and Humbert, 2002; Redman et al.,2015). However, the genetic differentiation between populationswas extremely low with only 0.006% of the genetic variation beingpartitioned between populations overall and only 5/66 pairwise Fstcomparisons showing statistically significant differentiation (high-est Fst value being 0.0351). This is consistent with previous studiesshowing genetic differentiation between H. contortus populationsto be relatively low in a number of different countries (includingthe UK, France and Sweden) (Silvestre et al., 2009; Redman et al.,2015). This has been suggested to be due to a combination of largeeffective population sizes and high gene flow by migration due toanthropogenic movement of livestock hosts (Blouin et al., 1995;Silvestre et al., 2009; Redman et al., 2015). The genetic differentia-tion in the populations examined here is lower than previouslyreported (Silvestre et al., 2009; Redman et al., 2015). For example,in our UK study using seven out of eight of the same markers usedhere, AMOVA analysis of microsatellite data revealed 2.84% of the

Page 8: International Journal for Parasitology...Genetic evidence for the spread of a benzimidazole resistance mutation across southern India from a single origin in the parasitic nematode

728 U. Chaudhry et al. / International Journal for Parasitology 45 (2015) 721–728

genetic variation was partitioned between populations overall(compared with 0.006% here). Also, in that study 10 out of 21 Fstcomparisons showed statistically significant differentiationbetween H. contortus populations (highest Fst was 0.0757) com-pared with just five out of 66 here (highest Fst was 0.0351). Thelower level of genetic differentiation between H. contortus popula-tions in southern India could be due to higher gene flow consistentwith the large amount of unregulated anthropogenic animalmovement that is believed to occur in the region (Raman et al.,2010; Swarnkar and Singh, 2010). This would predispose to therapid spread of anthelmintic resistance mutations across theregion. One caveat to this interpretation, however, is that the lowgenetic differentiation observed could also be at least partly dueto the high levels of larval survival on pasture in the year-roundwarm, humid climate which would minimise population bottle-necks compared with more temperate regions. A final point worthnoting is that there was no significant genetic differentiationdetected between the H. contortus populations taken from sheepand goats (Table 1). Indeed, the Fst values were lower betweenmany of the pairwise comparisons between sheep and goat popu-lations than between pairwise comparisons of these populations inthe same host. This supports the view that H. contortus is freelyshared between the two hosts with little or no host species barrier.

Acknowledgements

We are grateful to the Natural Sciences and EngineeringResearch Council of Canada (NSERC) (Grant numberRGPIN/371529-2209), NSERC-CREATE Host Pathogen Interactions(HPI) graduate training program at the University of Calgary,Canada as well as the BBSRC, United Kingdom for funding support(Grant number BB/E018505/1).

Appendix A. Supplementary data

Supplementary data associated with this article can be found, inthe online version, at http://dx.doi.org/10.1016/j.ijpara.2015.04.007.

References

Barrère, V., Alvarez, L., Suarez, G., Ceballos, L., Moreno, L., Lanusse, C., Prichard, R.K.,2012. Relationship between increased albendazole systemic exposure andchanges in single nucleotide polymorphisms on the b-tubulin isotype 1encoding gene in Haemonchus contortus. Vet. Parasitol. 186, 344–349.

Barrere, V., Falzon, L.C., Shakya, K.P., Menzies, P.I., Peregrine, A.S., Prichard, R.K.,2013a. Assessment of benzimidazole resistance in Haemonchus contortus insheep flocks in Ontario, Canada: comparison of detection methods for drugresistance. Vet. Parasitol. 198, 159–165.

Barrere, V., Keller, K., von Samson-Himmelstjerna, G., Prichard, R.K., 2013b.Efficiency of a genetic test to detect benzimidazole resistant Haemonchuscontortus nematodes in sheep farms in Quebec, Canada. Parasitol. Int. 62, 464–470.

Blouin, M.S., Yowell, C.A., Courtney, C.H., Dame, J.B., 1995. Host movement and thegenetic structure of populations of parasitic nematodes. Genetics 141, 1007–1014.

Brasil, B.S., Nunes, R.L., Bastianetto, E., Drummond, M.G., Carvalho, D.C., Leite, R.C.,Molento, M.B., Oliveira, D.A., 2012. Genetic diversity patterns of Haemonchusplacei and Haemonchus contortus populations isolated from domestic ruminantsin Brazil. Int. J. Parasitol. 42, 469–479.

Chandra, S., Prasad, A., Sankar, M., Yadav, N., Dalal, S., 2014. Molecular diagnosis ofbenzimidazole resistance in Haemonchus contortus in sheep from differentgeographic regions of North India. Vet. World 7.

Chaudhry, U., Miller, M., Yazwinski, T., Kaplan, R., Gilleard, J., 2014. The presence ofbenzimidazole resistance mutations in Haemonchus placei from US cattle. Vet.Parasitol. 204, 411–415.

Chaudhry, U., Redman, E., Abbas, M., Muthusamy, R., Ashraf, K., Gilleard, J., 2015.Genetic evidence for hybridisation between Haemonchus contortus andHaemonchus placei in natural field populations and its implications forinterspecies transmission of anthelmintic resistance. Int. J. Parasitol. 45, 149–159.

de Lourdes Mottier, M., Prichard, R.K., 2008. Genetic analysis of a relationshipbetween macrocyclic lactone and benzimidazole anthelmintic selection onHaemonchus contortus. Pharmacogenet. Genomics 18, 129–140.

Excoffier, L., Smouse, P.E., Quattro, J.M., 1992. Analysis of molecular varianceinferred from metric distances among DNA haplotypes: application to humanmitochondrial DNA restriction data. Genetics 131, 479–491.

Excoffier, L., Laval, G., Schneider, S., 2005. Arlequin (version 3.0): an integratedsoftware package for population genetics data analysis. Evol. Bioinform. Online1, 47–50.

Ghisi, M., Kaminsky, R., Maser, P., 2007. Phenotyping and genotyping of Haemonchuscontortus isolates reveals a new putative candidate mutation for benzimidazoleresistance in nematodes. Vet. Parasitol. 144, 313–320.

Grillo, V., Jackson, F., Cabaret, J., Gilleard, J.S., 2007. Population genetic analysis ofthe ovine parasitic nematode Teladorsagia circumcincta and evidence for acryptic species. Int. J. Parasitol. 37, 435–447.

Guo, S.W., Thompson, E.A., 1992. Performing the exact test of Hardy–Weinbergproportion for multiple alleles. Biometrics 48, 361–372.

Hoglund, J., Gustafsson, K., Ljungstrom, B.L., Engstrom, A., Donnan, A., Skuce, P.,2009. Anthelmintic resistance in Swedish sheep flocks based on a comparison ofthe results from the faecal egg count reduction test and resistant allelefrequencies of the beta-tubulin gene. Vet. Parasitol. 161, 60–68.

Hunt, P.W., Knox, M.R., Le Jambre, L.F., McNally, J., Anderson, L.J., 2008. Genetic andphenotypic differences between isolates of Haemonchus contortus in Australia.Int. J. Parasitol. 38, 885–900.

Huson, D.H., Bryant, D., 2006. Application of phylogenetic networks in evolutionarystudies. Mol. Biol. Evol. 23, 254–267.

Kotze, A.C., Cowling, K., Bagnall, N.H., Hines, B.M., Ruffell, A.P., Hunt, P.W., Coleman,G.T., 2012. Relative level of thiabendazole resistance associated with the E198Aand F200Y SNPs in larvae of a multi-drug resistant isolate of Haemonchuscontortus. Int. J. Parasitol. Drugs Drug Resist. 2, 92–97.

Laing, R., Kikuchi, T., Martinelli, A., Tsai, I.J., Beech, R.N., Redman, E., Holroyd, N.,Bartley, D.J., Beasley, H., Britton, C., Curran, D., Devaney, E., Gilabert, A., Hunt, M.,Jackson, F., Johnston, S.L., Kryukov, I., Li, K., Morrison, A.A., Reid, A.J., Sargison, N.,Saunders, G.I., Wasmuth, J.D., Wolstenholme, A., Berriman, M., Gilleard, J.S.,Cotton, J.A., 2013. The genome and transcriptome of Haemonchus contortus, akey model parasite for drug and vaccine discovery. Genome Biol. 14, R88.

Otsen, M., Plas, M.E., Groeneveld, J., Roos, M.H., Lenstra, J.A., Hoekstra, R., 2000.Genetic markers for the parasitic nematode Haemonchus contortus based onintron sequences. Exp. Parasitol. 95, 226–229.

Polzin, T., Daneschmand, S.V., 2003. On Steiner trees and minimum spanning treesin hypergraphs. Oper. Res. Lett. 31, 12–20.

Posada, D., 2008. JModelTest: phylogenetic model averaging. Mol. Biol. Evol. 25,1253–1256.

Raman, M., Selva, Bharathi R., Abdul, Basith S., Lalitha, J., 2010. Epizootiology ofgastrointestinal helminths in sheep and goats of Tamil Nadu. Indian J. SmallRumin. 16, 74–78.

Redman, E., Packard, E., Grillo, V., Smith, J., Jackson, F., Gilleard, J.S., 2008.Microsatellite analysis reveals marked genetic differentiation betweenHaemonchus contortus laboratory isolates and provides a rapid system ofgenetic fingerprinting. Int. J. Parasitol. 38, 111–122.

Redman, E., Whitelaw, F., Tait, A., Burgess, C., Bartley, Y., Skuce, P., Jackson, F.,Gilleard, J., 2015. The emergence of resistance to the benzimidazoleanthlemintics in parasitic nematodes of livestock is characterised by multipleindependent hard and soft selective sweeps. PLoS Negl. Trop. Dis. 9, e0003494.

Rice, W.R., 1989. Analysing tables of statistical tests. Evolution 43, 223–225.Rousset, F., Raymond, M., 1995. Testing heterozygote excess and deficiency.

Genetics 140, 1413–1419.Rufener, L., Kaminsky, R., Maser, P., 2009. In vitro selection of Haemonchus contortus

for benzimidazole resistance reveals a mutation at amino acid 198 of beta-tubulin. Mol. Biochem. Parasitol. 168, 120–122.

Silvestre, A., Cabaret, J., 2002. Mutation in position 167 of isotype 1 beta-tubulingene of Trichostrongylid nematodes: role in benzimidazole resistance?. Mol.Biochem. Parasitol. 120, 297–300.

Silvestre, A., Humbert, J.F., 2002. Diversity of benzimidazole-resistance alleles inpopulations of small ruminant parasites. Int. J. Parasitol. 32, 921–928.

Silvestre, A., Sauve, C., Cortet, J., Cabaret, J., 2009. Contrasting genetic structures of twoparasitic nematodes, determined on the basis of neutral microsatellite markersand selected anthelmintic resistance markers. Mol. Ecol. 18, 5086–5100.

Skuce, P., Stenhouse, L., Jackson, F., Hypsa, V., Gilleard, J., 2010. Benzimidazoleresistance allele haplotype diversity in United Kingdom isolates of Teladorsagiacircumcincta supports a hypothesis of multiple origins of resistance by recurrentmutation. Int. J. Parasitol. 40, 1247–1255.

Stevenson, L.A., Chilton, N.B., Gasser, R.B., 1995. Differentiation of Haemonchusplacei from H. contortus (Nematoda: Trichostrongylidae) by the ribosomal DNAsecond internal transcribed spacer. Int. J. Parasitol. 25, 483–488.

Swarnkar, C.P., Singh, D., 2010. Questionnaire survey on sheep husbandry and wormmanagement practices adopted by farmers in Rajasthan. Indian J. Small Rumin.16, 199–209.

Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M., Kumar, S., 2011. MEGA5:molecular evolutionary genetics analysis using maximum likelihood,evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol. 28,2731–2739.

Tiwari, J., Kumar, S., Kolte, A.P., Swarnkar, C.P., Singh, D., Pathak, K.M., 2006.Detection of benzimidazole resistance in Haemonchus contortus using RFLP-PCRtechnique. Vet. Parasitol. 138, 301–307, Epub 2006 Mar 29.

von Samson-Himmelstjerna, G., Walsh, T.K., Donnan, A.A., Carriere, S., Jackson, F.,Skuce, P.J., Rohn, K., Wolstenholme, A.J., 2009. Molecular detection ofbenzimidazole resistance in Haemonchus contortus using real-time PCR andpyrosequencing. Parasitology 136, 349–358.