Journal of General Virology (2016), 97, 747–755

DOI 10.1099/jgv.0.000378

Evidence of widespread natural recombination among field isolates of equine herpesvirus 4 but not among field isolates of equine herpesvirus 1 P. K. Vaz,1 J. Horsington,13 C. A. Hartley,1 G. F. Browning,1 N. P. Ficorilli,2 M. J. Studdert,2 J. R. Gilkerson2 and J. M. Devlin1 Correspondence Joanne Devlin

1

Asia-Pacific Centre for Animal Health, Faculty of Veterinary and Agricultural Sciences, The University of Melbourne, Parkville, Victoria 3010, Australia

[email protected]

2

Centre for Equine Infectious Diseases, Faculty of Veterinary and Agricultural Sciences, The University of Melbourne, Parkville, Victoria 3010, Australia

Received 12 August 2015 Accepted 17 December 2015

Recombination in alphaherpesviruses allows evolution to occur in viruses that have an otherwise stable DNA genome with a low rate of nucleotide substitution. High-throughput sequencing of complete viral genomes has recently allowed natural (field) recombination to be studied in a number of different alphaherpesviruses, however, such studies have not been applied to equine herpesvirus 1 (EHV-1) or equine herpesvirus 4 (EHV-4). These two equine alphaherpesviruses are genetically similar, but differ in their pathogenesis and epidemiology. Both cause economically significant disease in horse populations worldwide. This study used highthroughput sequencing to determine the full genome sequences of EHV-1 and EHV-4 isolates (11 and 14 isolates, respectively) from Australian or New Zealand horses. These sequences were then analysed and examined for evidence of recombination. Evidence of widespread recombination was detected in the genomes of the EHV-4 isolates. Only one potential recombination event was detected in the genomes of the EHV-1 isolates, even when the genomes from an additional 11 international EHV-1 isolates were analysed. The results from this study reveal another fundamental difference between the biology of EHV-1 and EHV-4. The results may also be used to help inform the future safe use of attenuated equine herpesvirus vaccines.

INTRODUCTION In many alphaherpesviruses, recombination is increasingly being recognized as an important mechanism that plays a major role in evolution of viruses with an otherwise stable DNA genome that have low rates of nucleotide substitution (Lee et al., 2013, 2012; Thiry et al., 2005). Recent advances in high-throughput sequencing of complete viral genomes have allowed natural (field) recombination to be studied in a number of different alphaherpesviruses 3Present address: Australian Animal Health Laboratory, CSIRO, 5 Portarlington Rd, East Geelong, Victoria 3220, Australia. The GenBank/EMBL/DDBJ accession numbers for the full EHV genome sequences determined in this study are KT324734, KT324724, KT324733, KT324732, KT324731, KT324730, KT324729, KT324728, KT324727, KT324726, KT324725, KT324748, KT324745, KT324738, KT324736, KT324743, KT324742, KT324740, KT324739, KT324737, KT324747, KT324741, KT324744, KT324735 and KT324746. Three supplementary figures and six supplementary tables are available with the online Supplementary Material.

000378 G 2015 The Authors

Printed in Great Britain

affecting animals and humans, including herpes simplex virus 1 (HSV-1), varicella-zoster virus (VZV) and infectious laryngotracheitis virus (ILTV) (Kolb et al., 2013; Lee et al., 2013; Norberg et al., 2015, 2006; Peters et al., 2006). Such studies provide insights into virus evolution and also allow the risk of recombination to be assessed in the context of attenuated vaccine use. In 2012, outbreaks of severe respiratory disease in poultry in Australia were attributed to natural recombination events between attenuated vaccine strains of ILTV that generated virulent recombinant viruses (Lee et al., 2012). These outbreaks of disease highlight the importance of studying and understanding herpesvirus recombination in order to protect animal health. Equine herpesvirus 1 (EHV-1) and equine herpesvirus 4 (EHV-4) are closely related alphaherpesviruses that cause economically significant disease in horses worldwide (Allen et al., 2004; Crabb & Studdert, 1995; Telford et al., 1992, 1998). Although EHV-1 and EHV-4 are genetically very similar, there are a number of important differences in their pathogenesis and epidemiology. Infection with 747

P. K. Vaz and others

EHV-4 is most commonly associated with upper respiratory tract disease, but is also occasionally associated with abortion (Allen et al., 2004; Patel & Heldens, 2005). Infection with EHV-1 also causes respiratory disease but infection frequently progresses beyond the upper respiratory tract to result in systemic disease, including abortion and myeloencephalitis (Allen et al., 2004; Edington et al., 1991; Patel & Heldens, 2005; Studdert et al., 2003). Sero-epidemiological studies have revealed a high prevalence of antibodies to EHV-4 in horse populations in different countries, including over 99 % sero-positivity in mares and foals tested on a large Thoroughbred stud farm in New South Wales, Australia (Gilkerson et al., 1999). Antibodies to EHV-1 are consistently detected at a lower prevalence than antibodies to EHV-4, with a large sero-epidemiological study in Australian horses detecting antibodies to EHV-1 in 26 % of mares and 11 % of foals (Gilkerson et al., 1999). Both EHV-1 and -4 have linear, double-stranded type ‘D’ DNA genomic structures that consist of a unique long (UL) and a unique short (US) genome region, with the US region flanked by large inverted repeats (internal repeat short, IRs, and terminal repeat short, TRs) (Telford et al., 1992, 1998). The EHV-1 and EHV-4 genomes are 150 kbp and 146 kbp in length, respectively. Both encode the same

76 homologous genes, with three duplicated genes in EHV-4 and four duplicated genes in EHV-1 within the repeat regions. The level of amino acid sequence identity between corresponding proteins encoded by the two genomes ranges from 55 % to 96 % (Telford et al., 1998). This study aimed to use high-throughput sequencing methods to determine the full genome sequences of a collection of diverse EHV-1 and EHV-4 isolates from Australian and New Zealand horses and to examine these sequences for evidence of recombination. We also aimed to assess the genetic diversity of the EHV-1 and EHV-4 isolates and to examine the phylogenetic relationships between the isolates.

RESULTS Complete genome sequences of 11 Australian EHV-1 isolates The full genome sequences of 11 Australian isolates of EHV-1 (Table 1) were determined by mapping against the reference sequence, or by de novo assembly. Sequence alignments from the former method of assembly are

Table 1. Australian and New Zealand equine herpesvirus isolates used in this study Virus ID EHV-1 438-77 NZA-77 717A-82 970-90 1029-93 1074-94 1966-02 2019-02 2222-03 3038-07 3045-07 EHV-4 ER39-67 R19-68 157-69 RT4-70 306-74 2387-75 405-76 3407-77 3409-77 475-78 960-90 1532-99 1546-99 3056-07 NA ,

Year

Region

Disease association

1977 1977 1982 1990 1993 1994 2002 2002 2003 2007 2007

Scone, NSW New Zealand VIC SA Wakefield, NSW Toowoomba, QLD Seymour, VIC Clayton, VIC Scone, NSW VIC NSW

Abortion Abortion Neurological/abortion Abortion Abortion Abortion Abortion Abortion Abortion Neurological Abortion

1967 1968 1969 1970 1974 1975 1976 1977 1977 1978 1990 1999 1999 2007

SA Toowoomba, QLD Yarra Glen, VIC Oakey, QLD Flemington, VIC Esperance, WA Flemington, VIC N/A Gosnells, WA Mt Pleasant, TAS Diggers Rest, VIC Flemington, VIC Toolern Vale, VIC TAS

Respiratory Respiratory Respiratory Respiratory Respiratory Abortion Respiratory Respiratory Abortion Abortion Respiratory Respiratory Respiratory Respiratory

Site of collection Fetus NA

Brain or fetus NA

Fetus Fetus Fetus Fetus Fetus Nasal septum NA

Respiratory tract Nasal swab Nasal swab Nasal swab Nasal swab Fetus Nasal swab NA

Fetus Fetus Nasal Nasal Nasal Nasal

swab swab swab swab

Reference

GenBank accession no.

Studdert & Blackney (1979) Horner (1981) Studdert et al. (1984) This study This study Varrasso et al. (2001) This study This study This study This study This study

KT324734 KT324724 KT324733 KT324732 KT324731 KT324730 KT324729 KT324728 KT324727 KT324726 KT324725

Studdert et al. (1970) Bagust & Pascoe (1968) Studdert (1971) Bagust & Pascoe (1970) Gleeson et al. (1976) Peet et al. (1978) Gleeson et al. (1976) This study Peet et al. (1978) Studdert & Blackney (1979) Varrasso et al. (2001) Huang Ja et al. (2002) Huang Ja et al. (2002) This study

KT324748 KT324745 KT324738 KT324736 KT324743 KT324742 KT324740 KT324739 KT324737 KT324747 KT324741 KT324744 KT324735 KT324746

Not available.

748

Journal of General Virology 97

Sequence and recombination analysis of EHV-1 and EHV-4

shown in Fig. 1. The results from de novo assembly were principally consistent with those produced by mapping against the reference sequence, with variation observed only in regions rich in repeats (Fig. S1, available in the online Supplementary Material). The estimated size of the EHV-1 genomes ranged from 148.37 kbp (isolate 2019-02) to 148.91 kbp (isolate 970-90). Sequence analysis of the EHV-1 genomes revealed a low degree of heterogeneity between isolates, with the most distant isolates (isolates 438-77, 3038-07, 2019-02 and 1966-02 compared to isolate 2222-03) sharing 99.9 % nucleotide identity, excluding large sequence gaps due to differences in short tandem repeat (STR) copy numbers. Details of the predicted amino acid differences between isolates are shown in Table S1. The number of single nucleotide polymorphisms (SNPs) between different isolates of EHV-1 is summarized in Table S2. The average number of SNPs between EHV-1 isolates was 96.

only in regions rich in repeats (Fig. S2). The estimated size of the EHV-4 full-length genomes ranged from 143.16 bp (isolate 3407-77) to 144.16 kbp (isolate 153299). Alignment of the 14 EHV-4 genomes revealed a number of sequence differences between isolates. The genetically most similar isolates shared 99.9 % nucleotide sequence identity (isolates ER39-67 and R19-68; isolates ER39-67 and 475-78; isolates 405-76 and 306-74), excluding large sequence gaps. The genetically most diverse isolates shared 98.9 % nucleotide identity (isolates 960-90 and 1532-99). Details of the predicted amino acid differences between isolates are shown in Table S3. Large regions of sequence difference (insertions/deletions, indels) were present in the repeat regions of the genome, between ORFs 64–65 and ORFs 66–67 of the IR/TR, in ORF24 of the UL region and in ORF71 in the US region. Smaller indels were seen in the UL region within ORFs 24 and 61 and in the US within ORFs 70, 71 and 75. The number of SNPs between different isolates of EHV-4 is summarized in Table S4. The average number of SNPs between EHV-4 isolates was 179.

Complete genome sequences of 14 Australian EHV-4 isolates The full genome sequences of 14 Australian isolates of EHV-4 (Table 1) were determined by mapping against the reference sequence, or by de novo assembly. Sequence alignments from the former method of assembly are shown in Fig. 2. The results from de novo assembly were principally consistent with those produced by mapping against the reference sequence, with variation observed

1

4,875

9,875

14,875

19,875

24,875

Recombination Phylogenetic recombination networks for EHV-1 and EHV-4 were generated from nucleotide alignments of the complete genome sequences, as well as alignments of the individual UL, US and IR regions, using SplitsTree4 (Huson, 1998). The multiple reticulate networks and

29,875

34,875

39,875

44,710

49,589

54,589

59,589

64,589

1. 438-77 UL

2. 2019-02 3. 1966-02 4. 970-90 5. 3038-07 6. 1074-94 7. 717A-82 8. 3045-07 9. NZA-77 10. 1029-93 11. 2222-03 69,588

74,524

79,524

84,524

89,524

94,524

99,524

104,524

109,488

114,049

119,048

123,842

128,842

136,911

1. 438-77 UL

IRS

US

2. 2019-02 3. 1966-02 4. 970-90 5. 3038-07 6. 1074-94 7. 717A-82 8. 3045-07 9. NZA-77 10. 1029-93 11. 2222-03

Fig. 1. Nucleotide sequence alignment for the complete genomes of Australasian EHV-1 isolates. Alignment of the complete genome sequences of EHV-1 isolates was performed using MAFFT . Our prototype strain 438-77 was used as the reference sequence. Vertical black lines indicate SNPs compared to the reference and dashes indicate sequence gaps. http://jgv.microbiologyresearch.org

749

P. K. Vaz and others

1

4,949

9,949

14,949

19,949

24,949

29,949

34,949

39,949

44,653

49,653

54,653

59,653

64,653

1. 405-76 UL

2. 306-74 3. 157-69 4. 3407-77 5. 3409-77 6. 2387-75 7. 960-90 8. 3056-07 9. 475-78 10. 1532-99 11. RT4-70 12. 1546-99 13. R19-68 14. ER39-67 69,653

74,593

79,593

84,593

89,593

94,593

99,592

104,592

109,547

114,385

119,370

124,117

128,912

134,979

1. 405-76 UL

IRS

US

2. 306-74 3. 157-69 4. 3407-77 5. 3409-77 6. 2387-75 7. 960-90 8. 3056-07 9. 475-78 10. 1532-99 11. RT4-70 12. 1546-99 13. R19-68 14. ER39-67

Fig. 2. Nucleotide sequence alignment for the complete genomes of Australasian EHV-4 isolates. Alignment of the complete genome sequences of EHV-4 isolates was performed using MAFFT . Our prototype strain EHV-4.405-76 was used as the reference sequence. Vertical black lines indicate SNPs compared to the reference and dashes indicate sequence gaps.

pair-wise homoplasy (Phi) test analyses (Fig. 3) indicate significant historical recombination events between the different EHV-4 isolates in the UL region (P50.001) but not in the IR or US regions. No significant recombination was detected between the 11 EHV-1 Australian/New Zealand isolates sequenced, irrespective of the genomic region analysed (Fig. 4). No significant recombination events were detected when the analyses were expanded to include the genome sequences from 11 additional international EHV-1 isolates (Fig S3). Detection of recombination breakpoints utilized six different methods: Recombination detection program (RDP), GENECONV , 3Seq, Maximum Chi Square (MaxChi), SiScan and BootScan. These methods were implemented in the program RDP4 (Martin et al., 2015). Analyses were performed using complete genome sequences, as well as alignments of the individual US, UL and IR regions. Consistent with the results from the SplitsTree analysis, evidence of potential recombination events was detected amongst EHV-4 isolates in the UL region using these different methods (Table 2). In addition, evidence of potential recombination events was detected in the US and IR regions of EHV-4 and another potential 750

recombination event was detected in the IR region of EHV-1 (Table 2). These potential recombination events had not been detected by SplitsTree and Phi test analyses.

DISCUSSION The 11 isolates of EHV-1 and 14 isolates of EHV-4 analysed in this study were all collected from Australian or New Zealand horses between 1967 and 2007. In Australia, EHV-4 was first isolated in 1967, whilst the first isolates of EHV-1 were not obtained until 1977, when it is believed EHV-1 was first introduced into Australia. The sequence data generated in this study have substantially increased the number of full genome sequences available for these viruses, particularly EHV-4, as only one EHV-4 genome has been fully sequenced previously (Telford et al., 1998). Evidence of extensive recombination was detected in the genomes of the EHV-4 isolates, but little or no recombination was detected between the genomes of the EHV-1 isolates, irrespective of whether only the genomes from Australian/New Zealand isolates of EHV-1 and -4 were analysed, or whether genomes from an additional 11 international EHV-1 isolates were included in the analysis. Journal of General Virology 97

Sequence and recombination analysis of EHV-1 and EHV-4

(b) Unique long

(a) Complete genome 960-90 2387-75

P = 1.665E-3

2387-75 960-90

3409-77 3407-77 157-69

P = 1.201E-3

3407-77 3409-77 157-69

405-76

405-76 306-74

306-74

3056-07

3056-07

0.1

0.1

RT4-70 1546-99 R19-68 ER39-67 1532-99 475-78

R19-68 1546-99 RT4-70 ER39-67 475-78 1532-99

(c) Unique short

(d) Repeat 960-90

P = 0.651

P = 1.0 2387-75

3056-07

ER39-67 475-78 1532-99 RT4-70 R19-68 1546-99

3409-77 3407-77 157-69 306-74 405-76

0.1 1546-99 R19-68 475-78 3056-07 ER39-67 RT4-70 1532-99

3409-77 3407-77 157-69 405-76 306-74

0.1 960-90 2387-75

Fig. 3. Recombination network trees generated from Australasian EHV-4 nucleotide alignments (excluding sequence repeats) using SplitsTree4. (a) Complete genome sequences, (b) UL region, (c) US region and (d) repeat region. The multiple reticulate networks indicate recombination events between the different isolates. The bar indicates the rate of evolution in sequence substitutions per site. The Phi test for detecting recombination, as implemented in SplitsTree4, was highly significant for the whole genome, and for the UL region, but not for the US region or the repeat region.

Analysis of the EHV-1 and EHV-4 full genome sequences showed that the genetic diversity amongst EHV-4 isolates was greater than the diversity among EHV-1 isolates. This is consistent with earlier studies that have assessed genetic differences between EHV-1 and EHV-4 isolates using other techniques, including sequencing of selected genes, or digestion with restriction endonucleases (Allen et al., 2004). This greater diversity in EHV-4 isolates, compared with EHV-1 isolates, is also consistent with the hypothesis that the progenitor virus of EHV-1 was transferred from another host to the ancestor of the modern horse more recently than EHV-4 (Crabb & Studdert, 1990) and has therefore had less time to diversify. An alternative hypothesis, arising from the results from this study, may be that the lack of genetic diversity within EHV-1 is a consequence of a low rate of http://jgv.microbiologyresearch.org

recombination in this virus species, given that recombination can contribute to genetic diversity. The lower level of genetic diversity in EHV-1, compared with EHV-4, would make recombination more difficult to detect in EHV-1. The total number of SNPs between any two EHV-1 isolates varied from 8 to 153 (excluding repeat sequences) with an average of 96 SNPs between any two isolates. It is possible that EHV-1 recombination events may have occurred but were not able to be detected at the required level of statistical significance, because an insufficient number of SNPs specific for the parental viruses were transferred to the recombinant viruses. However, evidence of frequent, natural recombination has been readily detected in VZV, which also has a highly conserved genome (approximately 30–200 SNP differences between 751

P. K. Vaz and others

(a) Complete genome

(b) Unique long 717A-82 717A-82

P = 0.690

P = 0.267

1074-94

1074-94

3045-07

3045-07 0.1

0.1

1029-93

1029-93

NZA-77

970-90

NZA-77

2019-02

970-90 1966-02 2222-03

1966-02

2019-02 438-77

3038-07

438-77

3038-07 2222-03

(c) Unique short

(d) Repeat 2222-03

1029-93 P = 1.0 P = 0.993

2222-03 0.1 1074-94

NZA-77

NZA-77

0.1 1029-93

717A-82

1074-94

3045-07 970-90 438-77 3038-07

1966-02

2019-02,970-90,1966-02

717A-82 3045-07

438-77 3038-07

2019-02

Fig. 4. Recombination network trees generated from Australasian EHV-1 nucleotide alignments (excluding sequence repeats) using SplitsTree4. (a) Complete genome sequences, (b) UL region, (c) US region and (d) repeat region. The bar indicates the rate of evolution in sequence substitutions per site. The Phi test for detecting recombination, as implemented in SplitsTree4, was not significant for the whole genome, or for any of the individual genome regions.

any two strains), using similar methods to those in this study (Norberg et al., 2015, 2006; Zell et al., 2012). Importantly, we assessed recombination in the EHV isolates using a number of different approaches. Simulated and empirical studies have demonstrated the value of using multiple methods for detecting recombination, rather than relying on a single approach (Posada, 2002; Posada & Crandall, 2001). The pair-wise homoplasy test (Phi test) implemented by SplitsTree is a robust and powerful method for detecting the presence/absence of recombination and has been shown to produce a low rate of false-positive results in a wide variety of circumstances (Bruen et al., 2006). This test detected recombination in the UL region of EHV-4, but did not detect recombination in EHV-1. Analysis of recombination breakpoints, using multiple methods within RDP4, produced some similar 752

results, but also detected evidence of other potential recombination events that were not detected by the Phi test, including within the US and IR regions of EHV-4 and one within the IR region of EHV-1. Although it is expected that different methods of detecting recombination will yield different results (Posada, 2002; Posada & Crandall, 2001), the consistent detection of multiple recombination events in EHV-4, but not in EHV-1, provides substantive evidence that recombination is widespread in EHV-4, but is limited or absent in EHV-1. The presence of widespread recombination in EHV-4, but not in EHV-1, may be due to differences in the epidemiology of the two viruses (Gilkerson et al., 1999), particularly the lower prevalence of EHV-1 infection and hence the lower rate of co-infection of the same animal with two or more EHV-1s, which is the fundamental requirement for Journal of General Virology 97

Sequence and recombination analysis of EHV-1 and EHV-4

Table 2. Recombination breakpoint analysis of EHV-1 and EHV-4 isolates Virus and region

Breakpoints* Breakpoint beginning 99 % CI

EHV-1 IR/TR EHV-4 UL UL UL IR/TR US

Possible viruses involved in recombination event

Method of breakpoint detectionD

Breakpoint ending 99 % CI

116122–120856

120865–123600

2019-02, 438-77, unknown

GENECONV ,

Undetermined (28) Undetermined (20689) Undetermined (96512)

28–6954 77657–98051 Undetermined (112236)

GENECONV , BootScan MaxChi, SiScan RDP, MaxChi, BootScan

Undetermined (121749) 122044–126553

Undetermined (122035) 126555–128536

306-74, 405-76, 475-78 3056-07, 475-78, RT4-70 475-78, 1546-99, 960-90, 3056-07, 306-74 3409-77, 3407-77, unknown 960-90, 2387-75, 3056-07, unknown

GENECONV , GENECONV ,

3Seq, BootScan

3Seq, BootScan, 3Seq, SiScan, BootScan

CI, Confidence interval. *Nucleotide number according to 438/77 (EHV-1) or 405/76 (EHV-4) reference sequences. DAlgorithm used to detect recombination, as implemented in RDP4.

intraspecific herpesvirus recombination (Thiry et al., 2005). Intrinsic virus-specific factors could also influence the propensity for recombination. Recent studies in HSV-1 have shown that the gene products of UL12 (59-to-39 exonuclease) and UL29 (ssDNA-binding protein) appear to function together as a two-component recombinase to promote recombination in infected cells (Weller & Sawitzke, 2014). The homologues of these genes in EHV-1 and EHV-4 (ORFs 50 and 31, respectively) have a high level of identity between the two virus species but their function has not been directly compared. Future studies to assess and compare recombination in EHV-1 and EHV-4 in cell culture are indicated. Natural recombination between herpesviruses has been recognized as a safety concern for live attenuated herpesvirus vaccines (Lee et al., 2012; Thiry et al., 2005). Although not currently used in Australia, attenuated vaccines are available for use, or are under development, in other countries to help control disease caused by infection with EHV-1 and EHV-4. The results from this study indicate that for attenuated EHV-4 vaccines, the risks of recombination should be considered when designing and implementing vaccination programmes, but that these considerations may be less of a concern for EHV-1 vaccines.

METHODS Viruses and cells used in this study. Eleven Australian or New

Zealand isolates of EHV-1 and 14 Australian isolates of EHV-4 (Table 1) were selected from our laboratory archives and propagated in cultures of equine fetal kidney cells (Studdert & Gleeson, 1977) for six passages, including three plaque purifications, in order to amplify sufficient virus for whole genome sequencing. Virions were purified by Ficoll gradient centrifugation (Lee et al., 2011). Alternatively, nucleocapsids were purified as described previously (Pignatti et al., 1979). Viral DNA was extracted from purified nucleocapsids using http://jgv.microbiologyresearch.org

standard phenol/chloroform extraction methods, or from purified virions using the High Pure PCR Template Preparation kit (Roche) according to manufacturer’s instructions. High-throughput sequencing and genome assembly. High-

throughput sequencing was performed as described previously (Lee et al., 2011) and the genome sequences for each isolate were assembled using Geneious V6.1.7 (Kearse et al., 2012), by mapping against the reference sequences of EHV-1 (GenBank accession NC_001491.2) or EHV-4 (GenBank accession NC_001844.1). An overview of the sequencing metrics (including mean coverage and read depth across each genome) is provided in Tables S5 (EHV-1) and S6 (EHV-4). The complete genome sequences of the EHV-1 and EHV-4 isolates were deposited in the NCBI GenBank database under the accession numbers listed in Table 1. For comparison, de novo assembly was also performed for two virus isolates (EHV-1 438-77 and EHV-4 405-76) using medium–low sensitivity settings in Geneious V6.1.7. Contig consensus sequences were aligned to the reference sequence to aid scaffold construction. Sequence alignment. Alignments of the complete genome

sequences (excluding the TR regions) of the EHV-1 and EHV-4 isolates, as well as the individual UL, US and IR regions, were prepared using the Multiple Alignment with Fast Fourier Transformation (MAFFT ) version 7 plugin in Geneious V6.1.7 (Katoh & Standley, 2013). The laboratory isolates 438-77 and 405-76 (Studdert & Blackney, 1979) were used as the reference sequences for EHV-1 and EHV-4 alignments, respectively. To expand the analysis of EHV-1 isolates, these same analyses were repeated after including publicly available full genome sequence data from an additional 11 EHV-1 isolates. These included five isolates from Japan (00c19, 90c16, 01c1, 89c105 and 89c25; GenBank accession numbers KF644576, KF644566, KF644578, KF644577 and KF644579, respectively), three isolates from the USA (FL06, NY05 and VA02; GenBank accession numbers KF644567, KF644570 and KF644572, respectively) and three isolates from the UK (NMKT04, V592 and Ab4; GenBank accession numbers KF644568, AY464052 and NC_001491.2, respectively). Recombination analyses. Tandem repeat regions were identified

and removed prior to recombination analysis. For this, perfect and 753

P. K. Vaz and others

imperfect repeats were identified in the aligned genomes using the Phobos plugin in Geneious V6.1.7 with default settings and score constraints for satellites. All repeat regions identified were deleted in all viruses in the alignment, including in those viruses in which the repeat region may not have been directly identified due to minor sequence variations that affected the repetitive nature of the sequence. In this way, the deletion of repeat regions did not change the alignment of the genomes. These alignments, and the isolated UL, US or internal repeat regions, were then used to detect evidence of intraspecific recombination using two recombination detection programs: SplitsTree4 (Huson, 1998) and RDP4 (Martin et al., 2015). Initially, only the genomes of the Australian/New Zealand EHV isolates were analysed. To expand upon the analyses of EHV1 isolates these same analyses were then repeated after including publicly available full genome sequence data from the additional 11 EHV-1 isolates described above.

the Hunter Valley of New South Wales between 1995 and 1997. Vet Microbiol 68, 15–25.

In SplitsTree4, splits network trees were generated with an uncorrected P characters transformation model, ignoring constant sites. Other models were tested but resulted in no significant topological differences in the generated trees. Statistical analyses of the recombination networks were performed using the Phi test (Bruen et al., 2006) as implemented by SplitsTree4. In RDP4, six different methods were used to assess the sequences for recombination breakpoints: RDP, GENECONV , 3Seq, SiScan, MaxChi and BootScan. Default RDP4 settings were used throughout. Only breakpoints with a Bonferroni-corrected P value v0.05 were reported. Duplicate breakpoints were omitted.

Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock, S., Buxton, S., Cooper, A., Markowitz, S. & other authors (2012). Geneious Basic: an integrated and extendable desktop

Gleeson, L. J., Sullivan, N. D. & Studdert, M. J. (1976). Equine

herpesviruses: type 3 as an abortigenic agent. Aust Vet J 52, 349–354. Horner, G. W. (1981). Serological relationship between abortifacient

and respiratory strains of equine herpesvirus type 1 in New Zealand. N Z Vet J 29, 7–8. Huang Ja, J. A., Ficorilli, N., Hartley, C. A., Allen, G. P. & Studdert, M. J. (2002). Polymorphism of open reading frame 71 of equine

herpesvirus-4 (EHV-4) and EHV-1. J Gen Virol 83, 525–531. H. (1998). SplitsTree: analyzing and visualizing evolutionary data. Bioinformatics 14, 68–73.

Huson, D.

MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol 30, 772–780.

Katoh, K. & Standley, D. M. (2013).

software platform for the organization and analysis of sequence data. Bioinformatics 28, 1647–1649. Kolb, A. W., Ane´, C. & Brandt, C. R. (2013). Using HSV-1 genome

phylogenetics to track past human migrations. PLoS One 8, e76267. Lee, S. W., Markham, P. F., Markham, J. F., Petermann, I., Noormohammadi, A. H., Browning, G. F., Ficorilli, N. P., Hartley, C. A. & Devlin, J. M. (2011). First complete genome sequence of

infectious laryngotracheitis virus. BMC Genomics 12, 197.

ACKNOWLEDGEMENTS

Lee, S. W., Markham, P. F., Coppo, M. J., Legione, A. R., Markham, J. F., Noormohammadi, A. H., Browning, G. F., Ficorilli, N., Hartley, C. A. & Devlin, J. M. (2012). Attenuated vaccines can recombine to

This work was supported by the Australian Research Council (ARC, DP130103991). J. M. D. is supported by an ARC Future Fellowship (FT140101287). The authors thank J. Cuxson for her assistance.

Lee, S. W., Devlin, J. M., Markham, J. F., Noormohammadi, A. H., Browning, G. F., Ficorilli, N. P., Hartley, C. A. & Markham, P. F. (2013). Phylogenetic and molecular epidemiological studies reveal

REFERENCES Allen, G., Kydd, J. & Slater, J. (2004). Equid herpesvirus 1 and

equid herpesvirus 4 infections. In Infectious Diseases of Livestock, pp. 829–859. Edited by J. Coetzer & R. Tustin. Newmarket: Oxford University Press. Bagust, T. J. & Pascoe, R. R. (1968). Isolation of equine

rhinopnemonitis virus from acute respiratory disease in a horse in queensland. Vet J 44, 296. Bagust, T. J. & Pascoe, R. R. (1970). Studies on equine herpesviruses.

1. Characterisation of a strain of equine rhinopneumonitis virus isolated in Queensland. Aust Vet J 46, 421–427.

form virulent field viruses. Science 337, 188.

evidence of multiple past recombination events between infectious laryngotracheitis viruses. PLoS One 8, e55121. Martin, D., Murrell, B., Golden, M., Khoosal, A. & Muhire, B. (2015).

RDP4: detection and analysis of recombination patterns in virus genomes. Virus Evolution 1, vev003. Norberg, P., Liljeqvist, J. A., Bergstro¨m, T., Sammons, S., Schmid, D. S. & Loparev, V. N. (2006). Complete-genome phylogenetic

approach to varicella-zoster virus evolution: genetic divergence and evidence for recombination. J Virol 80, 9569–9576. Norberg, P., Depledge, D. P., Kundu, S., Atkinson, C., Brown, J., Haque, T., Hussaini, Y., MacMahon, E., Molyneaux, P. & other authors (2015). Recombination of globally circulating varicella-

zoster virus. J Virol 89, 7133–7146.

statistical test for detecting the presence of recombination. Genetics 172, 2665–2681.

Patel, J. R. & Heldens, J. (2005). Equine herpesviruses 1 (EHV-1) and 4 (EHV-4) – epidemiology, disease and immunoprophylaxis: a brief review. Vet J 170, 14–23.

Crabb, B. S. & Studdert, M. J. (1990). Comparative studies of the

Peet, R. L., Coackley, W., Smith, V. W. & Main, C. (1978). Equine

Bruen, T. C., Philippe, H. & Bryant, D. (2006). A simple and robust

proteins of equine herpesviruses 4 and 1 and asinine herpesvirus 3: antibody response of the natural hosts. J Gen Virol 71, 2033–2041. Crabb, B. S. & Studdert, M. J. (1995). Equine herpesviruses 4 (equine

rhinopneumonitis virus) and 1 (equine abortion virus). Adv Virus Res 45, 153–190. Edington, N., Smyth, B. & Griffiths, L. (1991). The role of endothelial

cell infection in the endometrium, placenta and foetus of equid herpesvirus 1 (EHV-1) abortions. J Comp Pathol 104, 379–387. Gilkerson, J. R., Whalley, J. M., Drummer, H. E., Studdert, M. J. & Love, D. N. (1999). Epidemiological studies of equine herpesvirus 1

(EHV-1) in Thoroughbred foals: a review of studies conducted in 754

abortion associated with herpesvirus. Aust Vet J 54, 151. Peters, G. A., Tyler, S. D., Grose, C., Severini, A., Gray, M. J., Upton, C. & Tipples, G. A. (2006). A full-genome phylogenetic analysis of

varicella-zoster virus reveals a novel origin of replication-based genotyping scheme and evidence of recombination between major circulating clades. J Virol 80, 9850–9860. Pignatti, P. F., Cassai, E., Meneguzzi, G., Chenciner, N. & Milanesi, G. (1979). Herpes simplex virus DNA isolation from infected cells with a

novel procedure. Virology 93, 260–264. Posada, D. (2002). Evaluation of methods for detecting recombination from DNA sequences: empirical data. Mol Biol Evol 19, 708–717.

Journal of General Virology 97

Sequence and recombination analysis of EHV-1 and EHV-4

Posada, D. & Crandall, K. A. (2001). Evaluation of methods for

computer

Outbreak of equine herpesvirus type 1 myeloencephalitis: new insights from virus identification by PCR and the application of an EHV-1-specific antibody detection ELISA. Vet Rec 153, 417–423.

Studdert, M. J. (1971). Equine herpesviruses. 4. Concurrent infection

Telford, E. A., Watson, M. S., McBride, K. & Davison, A. J. (1992). The

detecting recombination from DNA sequences: simulations. Proc Natl Acad Sci U S A 98, 13757–13762.

in horses with strangles and conjunctivitis. Aust Vet J 47, 434–436.

DNA sequence of equine herpesvirus-1. Virology 189, 304–316.

Studdert, M. J. & Blackney, M. H. (1979). Equine herpesviruses: on the

Telford, E. A., Watson, M. S., Perry, J., Cullinane, A. A. & Davison, A. J. (1998). The DNA sequence of equine herpesvirus-4. J Gen Virol 79,

differentiation of respiratory from foetal strains of type 1. Aust Vet J 55, 488–492. Studdert, M. J. & Gleeson, L. J. (1977). Isolation of equine rhinovirus

1197–1203.

type 1. Aust Vet J 53, 452.

Thiry, E., Meurens, F., Muylkens, B., McVoy, M., Gogev, S., Thiry, J., Vanderplasschen, A., Epstein, A., Keil, G. & Schynts, F. (2005).

Studdert, M. J., Turner, A. J. & Peterson, J. E. (1970). Equine

Recombination in alphaherpesviruses. Rev Med Virol 15, 89–103.

herpesviruses. I. Isolation and characterisation of equine rhinopneumonitis virus and other equine herpesviruses from horses. Aust Vet J 46, 83–89.

Varrasso, A., Dynon, K., Ficorilli, N., Hartley, C. A., Studdert, M. J. & Drummer, H. E. (2001). Identification of equine herpesviruses 1 and

Studdert, M. J., Fitzpatrick, D. R., Horner, G. W., Westbury, H. A. & Gleeson, L. J. (1984). Molecular epidemiology and pathogenesis

Weller, S. K. & Sawitzke, J. A. (2014). Recombination promoted by

4 by polymerase chain reaction. Aust Vet J 79, 563–569. DNA viruses: phage l to herpes simplex virus. Annu Rev Microbiol 68, 237–258.

of some equine herpesvirus type 1 (equine abortion virus) and type 4 (equine rhinopneumonitis virus) isolates. Aust Vet J 61, 345–348.

Zell, R., Taudien, S., Pfaff, F., Wutzler, P., Platzer, M. & Sauerbrei, A. (2012). Sequencing of 21 varicella-zoster virus genomes reveals two

Studdert, M. J., Hartley, C. A., Dynon, K., Sandy, J. R., Slocombe, R. F., Charles, J. A., Milne, M. E., Clarke, A. F. & El-Hage, C. (2003).

novel genotypes and evidence of recombination. J Virol 86, 1608–1622.

http://jgv.microbiologyresearch.org

755

Evidence of widespread natural recombination among field isolates of equine herpesvirus 4 but not among field isolates of equine herpesvirus 1.

Recombination in alphaherpesviruses allows evolution to occur in viruses that have an otherwise stable DNA genome with a low rate of nucleotide substi...
NAN Sizes 0 Downloads 7 Views