Download Evidence of widespread natural recombination among field isolates

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Orthohantavirus wikipedia , lookup

Hepatitis B wikipedia , lookup

Marburg virus disease wikipedia , lookup

Middle East respiratory syndrome wikipedia , lookup

Influenza A virus wikipedia , lookup

HIV wikipedia , lookup

Henipavirus wikipedia , lookup

Neisseria meningitidis wikipedia , lookup

Herpes simplex virus wikipedia , lookup

Transcript
1
Evidence of widespread natural recombination among field isolates of equine
2
herpesvirus 4 but not among field isolates of equine herpesvirus 1
3
4
Running title: Sequence and recombination analysis of EHV-1 and EHV-4
5
6
P.K. Vaz,1 J. Horsington,1,2 C.A. Hartley,1 G.F. Browning,1 N. P. Ficorilli,3 M.J. Studdert,3 J.R.
7
Gilkerson,3 and J.M Devlin1
8
9
1Asia-Pacific
Centre for Animal Health, Faculty of Veterinary and Agricultural Sciences, The
10
University of Melbourne, Parkville, Victoria, 3010, Australia.
11
2
12
Victoria, 3220, Australia.
13
3 Centre
14
University of Melbourne, Parkville, Victoria, 3010, Australia.
15
Author for correspondence: Joanne Devlin, ph: +61 3 9035 8110, fax: +61 3 8344 7374, email:
16
[email protected]
Present address: Australian Animal Health Laboratory, CSIRO, 5 Portarlington Rd, East Geelong,
for Equine Infectious Diseases, Faculty of Veterinary and Agricultural Sciences, The
17
18
Main text = 2788 words. Summary = 196 words. Tables = 2. Figures = 4.
19
20
The GenBank accession numbers for the full EHV genome sequences determined in this study are
21
KT324734, KT324724, KT324733, KT324732, KT324731, KT324730, KT324729, KT324728, KT324727,
22
KT324726, KT324725, KT324748, KT324745, KT324738, KT324736, KT324743, KT324742, KT324740,
23
KT324739, KT324737, KT324747, KT324741, KT324744, KT324735 and KT324746.
1
24
SUMMARY
25
Recombination in alphaherpesviruses allows evolution to occur in viruses that have an otherwise
26
stable DNA genome with a low rate of nucleotide substitution. High-throughput sequencing of
27
complete viral genomes has recently allowed natural (field) recombination to be studied in a
28
number of different alphaherpesviruses, however such studies have not been applied to equine
29
herpesvirus 1 (EHV-1) or equine herpesvirus 4 (EHV-4). These two equine alphaherpesviruses are
30
genetically similar, but differ in their pathogenesis and epidemiology. Both cause economically
31
significant disease in horse populations worldwide. This study used high-throughput sequencing to
32
determine the full genome sequences of EHV-1 and EHV-4 isolates (11 and 14 isolates, respectively)
33
from Australian or New Zealand horses. These sequences were then analysed and examined for
34
evidence of recombination. Evidence of widespread recombination was detected in the genomes of
35
the EHV-4 isolates. Only one potential recombination event was detected in the genomes of the
36
EHV-1 isolates, even when the genomes from an additional 11 international EHV-1 isolates were
37
analysed. The results from this study reveal another fundamental difference between the biology of
38
EHV-1 and EHV-4. The results may also be used to help inform the future safe use of attenuated
39
equine herpesvirus vaccines.
40
2
41
INTRODUCTION
42
In many alphaherpesviruses, recombination is increasingly being recognised as an important
43
mechanism that plays a major role in evolution of viruses with an otherwise stable DNA genome
44
that have low rates of nucleotide substitution (Lee et al., 2013; Lee et al., 2012; Thiry et al., 2005).
45
Recent advances in high-throughput sequencing of complete viral genomes has allowed natural
46
(field) recombination to be studied in a number of different alphaherpesviruses affecting animals
47
and humans, including herpes simplex virus 1 (HSV-1), varicella-zoster virus (VZV) and infectious
48
laryngotracheitis virus (ILTV) (Kolb et al., 2013; Lee et al., 2013; Norberg et al., 2015; Norberg et al.,
49
2006; Peters et al., 2006). Such studies provide insights into virus evolution and also allow the risk
50
of recombination to be assessed in the context of attenuated vaccine use. In 2012, outbreaks of
51
severe respiratory disease in poultry in Australia were attributed to natural recombination events
52
between attenuated vaccine strains of ILTV that generated virulent recombinant viruses (Lee et al.,
53
2012). These outbreaks of disease highlight the importance of studying and understanding
54
herpesvirus recombination in order to protect animal health.
55
Equine herpesvirus 1 (EHV-1) and equine herpesvirus 4 (EHV-4) are closely related
56
alphaherpesviruses that cause economically significant disease in horses worldwide (Allen et al.,
57
2004; Crabb & Studdert, 1995; Telford et al., 1992; Telford et al., 1998). Although EHV-1 and EHV-4
58
are genetically very similar, there are a number of important differences in their pathogenesis and
59
epidemiology. Infection with EHV-4 is most commonly associated with upper respiratory tract
60
disease, but is also occasionally associated with abortion (Allen et al., 2004; Patel & Heldens, 2005).
61
Infection with EHV-1 also causes respiratory disease but infection frequently progresses beyond the
62
upper respiratory tract to result in systemic disease, including abortion and myeloencephalitis (Allen
63
et al., 2004; Edington et al., 1991; Patel & Heldens, 2005; Studdert et al., 2003). Sero-
3
64
epidemiological studies have revealed a high prevalence of antibodies to EHV-4 in horse
65
populations in different countries, including over 99% sero-positivity in mares and foals tested on a
66
large Thoroughbred stud farm in New South Wales, Australia (Gilkerson et al., 1999). Antibodies to
67
EHV-1 are consistently detected at a lower prevalence than antibodies to EHV-4, with a large sero-
68
epidemiological study in Australian horses detecting antibodies to EHV-1 in 26% of mares and 11%
69
of foals (Gilkerson et al., 1999).
70
Both EHV-1 and -4 have linear, double-stranded type ‘D’ DNA genomic structures that consist of a
71
unique long (UL) and a unique short (US) genome region, with the US region flanked by large
72
inverted repeats (internal repeat short, IRs, and terminal repeat short, TRs) (Telford et al., 1992;
73
Telford et al., 1998). The EHV-1 and EHV-4 genomes are 150 kbp and the 146 kbp in length,
74
respectively. Both encode the same 76 homologous genes, with three duplicated genes in EHV-4
75
and four duplicated genes in EHV-1 within the repeat regions. The level of amino acid sequence
76
identity between corresponding proteins encoded by the two genomes ranges from 55% to 96%
77
(Telford et al., 1998).
78
This study aimed to use high-throughput sequencing methods to determine the full genome
79
sequences of a collection of diverse EHV-1 and EHV-4 isolates from Australian and New Zealand
80
horses and to examine these sequences for evidence of recombination. We also aimed to assess the
81
genetic diversity of the EHV-1 and EHV-4 isolates and to examine the phylogenetic relationships
82
between the isolates.
83
84
RESULTS
85
86
Complete genome sequences of 11 Australian EHV-1 isolates
4
87
The full genome sequences of 11 Australian isolates of EHV-1 (Table 1) were determined by
88
mapping against the reference sequence, or by de novo assembly. Sequence alignments from the
89
former method of assembly are shown in Fig. 1. The results from de novo assembly were principally
90
consistent with those produced by mapping against the reference sequence, with variation
91
observed only in regions rich in repeats (Fig. S1). The estimated size of the EHV-1 genomes ranged
92
from 148.37 kbp (isolate 2019-02) to 148.91 kbp (isolate 970-90). Sequence analysis of the EHV-1
93
genomes revealed a low degree of heterogeneity between isolates, with the most distant isolates
94
(isolates 438-77, 3038-07, 2019-02 and 1966-02 compared to isolate 2222-03) sharing 99.9%
95
nucleotide identity, excluding large sequence gaps due to differences in short tandem repeat (STR)
96
copy numbers. Details of the predicted amino acid differences between isolates are shown in Table
97
S1. The number of single nucleotide polymorphisms (SNPs) between different isolates of EHV-1 is
98
summarised in Table S2. The average number of SNPs between EHV-1 isolates was 96.
99
100
Complete genome sequences of 14 Australian EHV-4 isolates
101
The full genome sequences of 14 Australian isolates of EHV-4 (Table 1) were determined by
102
mapping against the reference sequence, or by de novo assembly. Sequence alignments from the
103
former method of assembly are shown in Fig. 2. The results from de novo assembly were principally
104
consistent with those produced by mapping against the reference sequence, with variation
105
observed only in regions rich in repeats (Fig. S2). The estimated size of the EHV-4 full-length
106
genomes ranged from 143.16 bp (isolate 3407-77) to 144.16 kbp (isolate 1532-99). Alignment of the
107
14 EHV-4 genomes revealed a number of sequence differences between isolates. The genetically
108
most similar isolates shared 99.9% nucleotide sequence identity (isolates ER39-67 and R19-68;
109
isolates ER39-67 and 475-78; isolates 405-76 and 306-74), excluding large sequence gaps. The
110
genetically most diverse isolates shared 98.9% nucleotide identity (isolates 960-90 and 1532-99).
5
111
Details of the predicted amino acid differences between isolates are shown in Table S3. Large
112
regions of sequence difference (insertions/deletions, indels) were present in the repeat regions of
113
the genome, between ORFs 64-65 and ORFs 66-67 of the IR/TR, in ORF24 of the UL region, and in
114
ORF71 in the US region. Smaller indels were seen in the UL region within ORFs 24 and 61, and in the
115
US within ORFs 70, 71 and 75. The number of SNPs between different isolates of EHV-4 is
116
summarised in Table S4. The average number of SNPs between EHV-4 isolates was 179.
117
Recombination
118
Phylogenetic recombination networks for EHV-1 and EHV-4 were generated from nucleotide
119
alignments of the complete genome sequences, as well alignments of the individual UL, US and IR
120
regions, using SplitsTree4 (Huson, 1998). The multiple reticulate networks and pair-wise homoplasy
121
(Phi) test analyses (Fig. 3) indicate significant historical recombination events between the different
122
EHV-4 isolates in the UL region (P = 0.001) but not in the IR or US regions. No significant
123
recombination was detected between the 11 EHV-1 Australian/New Zealand isolates sequenced,
124
irrespective of the genomic region analysed (Fig. 4). No significant recombination events were
125
detected when the analyses were expanded to include the genome sequences from 11 additional
126
international EHV-1 isolates (Fig S3).
127
Detection of recombination breakpoints utilised six different methods: Recombination detection
128
program (RDP), GENECONV, 3Seq, Maximum Chi Square (MaxChi), SiScan and Bootscan. These
129
methods were implemented in the program RDP4 (Martin et al., 2015). Analyses were performed
130
using complete genome sequences, as well as alignments of the individual US, UL and IR regions.
131
Consistent with the results from the SplitsTree analysis, evidence of potential recombination events
132
was detected amongst EHV-4 isolates in the UL region using these different methods (Table 2). In
133
addition, evidence of potential recombination events was detected in the US and IR regions of EHV-
6
134
4, and another potential recombination event was detected in the IR region of EHV-1 (Table 2).
135
These potential recombination events had not been detected by SplitsTree and Phi test analyses.
136
137
DISCUSSION
138
The 11 isolates of EHV-1 and 14 isolates of EHV-4 analysed in this study were all collected from
139
Australian or New Zealand horses between 1967 and 2007. In Australia, EHV-4 was first isolated in
140
1967, whilst the first isolates of EHV-1 were not obtained until 1977, when it is believed EHV-1 was
141
first introduced into Australia. The sequence data generated in this study has substantially
142
increased the number of full genome sequences available for these viruses, particularly EHV-4, as
143
only one EHV-4 genome has been fully sequenced previously (Telford et al., 1998). Evidence of
144
extensive recombination was detected in the genomes of the EHV-4 isolates, but little or no
145
recombination was detected between the genomes of the EHV-1 isolates, irrespective of whether
146
only the genomes from Australian/New Zealand isolates of EHV-1 and -4 were analysed, or whether
147
genomes from an additional 11 international EHV-1 isolates were included in the analysis.
148
Analysis of the EHV-1 and EHV-4 full genome sequences showed that the genetic diversity amongst
149
EHV-4 isolates was greater than the diversity among EHV-1 isolates. This is consistent with earlier
150
studies that have assessed genetic differences between EHV-1 and EHV-4 isolates using other
151
techniques, including sequencing of selected genes, or digestion with restriction endonucleases
152
(Allen et al., 2004). This greater diversity in EHV-4 isolates, compared with EHV-1 isolates, is also
153
consistent with the hypothesis that the progenitor virus of EHV-1 was transferred from another host
154
to the ancestor of the modern horse more recently than EHV-4 (Crabb & Studdert, 1990) and has
155
therefore had less time to diversify. An alternative hypothesis, arising from the results from this
156
study, may be that the lack of genetic diversity within EHV-1 is a consequence of a low rate of
157
recombination in this virus species, given that recombination can contribute to genetic diversity.
7
158
The lower level of genetic diversity in EHV-1, compared to EHV-4, would make recombination more
159
difficult to detect in EHV-1. The total number of SNPs between any two EHV-1 isolates varied from 8
160
to 153 (excluding repeat sequences) with an average of 96 SNPs between any two isolates. It is
161
possible that EHV-1 recombination events may have occurred but were not able to be detected at
162
the required level of statistical significance because an insufficient number of SNPs specific for the
163
parental viruses were transferred to the recombinant viruses. However, evidence of frequent,
164
natural recombination has been readily detected in VZV, which also has a highly conserved genome
165
(approximately 30 – 200 SNP differences between any two strains) using similar methods to those in
166
this study (Norberg et al., 2015; Norberg et al., 2006; Zell et al., 2012).
167
Importantly we assessed recombination in the EHV isolates using a number of different approaches.
168
Simulated and empirical studies have demonstrated the value of using multiple methods for
169
detecting recombination, rather than relying on a single approach (Posada, 2002; Posada &
170
Crandall, 2001). The pair-wise homoplasy test (Phi test) implemented by SplitsTree is a robust and
171
powerful method for detecting the presence/absence of recombination and has been shown to
172
produce a low rate of false positive results in a wide variety of circumstances (Bruen et al., 2006).
173
This test detected recombination in the UL region of EHV-4, but did not detect recombination in
174
EHV-1. Analysis of recombination breakpoints, using multiple methods within RDP4, produced some
175
similar results, but also detected evidence of other potential recombination events that were not
176
detected by the Phi test, including within the US and IR regions of EHV-4 and one within the IR
177
region of EHV-1. Although it is expected that different methods of detecting recombination will
178
yield different results (Posada, 2002; Posada & Crandall, 2001) the consistent detection of multiple
179
recombination events in EHV-4, but not in EHV-1, provides substantive evidence that recombination
180
is widespread in EHV-4, but is limited or absent in EHV-1.
8
181
The presence of widespread recombination in EHV-4, but not in EHV-1, may be due to differences in
182
the epidemiology of the two viruses (Gilkerson et al., 1999), particularly the lower prevalence of
183
EHV-1 infection and hence the lower rate of co-infection of the same animal with two or more EHV-
184
1s, which is the fundamental requirement for intraspecific herpesvirus recombination (Thiry et al.,
185
2005). Intrinsic virus-specific factors could also influence the propensity for recombination. Recent
186
studies in HSV-1 have shown that the gene products of UL12 (5’-to-3’ exonuclease) and UL29 (single
187
stranded DNA binding protein) appear to function together as a two-component recombinase to
188
promote recombination in infected cells (Weller & Sawitzke, 2014). The homologues of these genes
189
in EHV-1 and EHV-4 (ORFs 50 and 31, respectively) have a high level of identity between the two
190
virus species but their function has not been directly compared. Future studies to assess and
191
compare recombination in EHV-1 and EHV-4 in cell culture are indicated.
192
Natural recombination between herpesviruses has been recognised as a safety concern for live
193
attenuated herpesvirus vaccines (Lee et al., 2012; Thiry et al., 2005). Although not currently used in
194
Australia, attenuated vaccines are available for use, or are under development, in other countries to
195
help control disease caused by infection with EHV-1 and EHV-4. The results from this study indicate
196
that for attenuated EHV-4 vaccines the risks of recombination should be considered when designing
197
and implementing vaccination programs, but that these considerations may be less of a concern for
198
EHV-1 vaccines.
199
METHODS
200
Viruses and cells used in this study
201
Eleven Australian or New Zealand isolates of EHV-1 and 14 Australian isolates of EHV-4 (Table 1)
202
were selected from our laboratory archives and propagated in cultures of equine foetal kidney cells
203
(Studdert & Gleeson, 1977) for six passages, including three plaque purifications, in order to amplify
9
204
sufficient virus for whole genome sequencing. Virions were purified by Ficoll gradient centrifugation
205
(Lee et al., 2011). Alternatively, nucleocapsids were purified as described previously (Pignatti et al.,
206
1979). Viral DNA was extracted from purified nucleocapsids using standard phenol-chloroform
207
extraction methods, or from purified virions using the High Pure PCR Template Preparation Kit
208
(Roche) according to manufacturer’s instructions.
209
210
High throughput sequencing and genome assembly
211
High-throughput sequencing was performed as described previously (Lee et al., 2011) and the
212
genome sequences for each isolate were assembled using Geneious V6.1.7 (Kearse et al., 2012) by
213
mapping against the reference sequences of EHV-1 (GenBank accession NC_001491.2) or EHV-4
214
(GenBank accession NC_001844.1). An overview of the sequencing metrics (including mean
215
coverage and read depth across each genome) is provided in Tables S5 (EHV-1) and S6 (EHV-4). The
216
complete genome sequences of the EHV-1 and EHV-4 isolates were deposited in the NCBI GenBank
217
database under the accession numbers listed in Table 1. For comparison, de novo assembly was also
218
performed for two virus isolates (EHV-1 438-77 and EHV-4 405-76) using medium-low sensitivity
219
settings in Geneious V6.1.7. Contig consensus sequences were aligned to the reference sequence to
220
aid scaffold construction.
221
222
Sequence alignment
223
Alignments of the complete genome sequences (excluding the TR regions) of the EHV-1 and EHV-4
224
isolates, as well as the individual UL, US and IR regions, were prepared using the Multiple Alignment
225
with Fast Fourier Transformation (MAFFT) version 7 plugin in Geneious V6.1.7 (Katoh & Standley,
226
2013). The laboratory isolates 438-77 and 405-76 (Studdert & Blackney, 1979) were used as the
227
reference sequences for EHV-1 and EHV-4 alignments, respectively.
10
228
To expand the analysis of EHV-1 isolates, these same analyses were repeated after including
229
publically available full genome sequence data from an additional 11 EHV-1 isolates. These included
230
five isolates from Japan (00c19, 90c16, 01c1, 89c105 and 89c25; GenBank accession numbers
231
KF644576, KF644566, KF644578, KF644577 and KF644579, respectively), three isolates from the
232
USA (FL06, NY05 and VA02; GenBank accession numbers KF644567, KF644570 and KF644572,
233
respectively) and three isolates from the UK (NMKT04, V592 and Ab4; GenBank accession numbers
234
KF644568, AY464052 and NC_001491.2, respectively).
235
236
Recombination analyses
237
Tandem repeat regions were identified and removed prior to recombination analysis. For this,
238
perfect and imperfect repeats were identified in the aligned genomes using the Phobos plugin in
239
Geneious V6.1.7 with default settings and score constraints for satellites. All repeat regions
240
identified were deleted in all viruses in the alignment, including in those viruses in which the repeat
241
region may not have been directly identified due to minor sequence variations that affected the
242
repetitive nature of the sequence. In this way, the deletion of repeat regions did not change the
243
alignment of the genomes. These alignments, and the isolated UL, US or internal repeat regions,
244
were then used to detect evidence of intraspecific recombination using two recombination
245
detection programs; SplitsTree4 (Huson, 1998) and RDP4 (Martin et al., 2015). Initially only the
246
genomes of the Australian/New Zealand EHV isolates were analyzed. To expand upon the analyses
247
of EHV1 isolates these same analyses were then repeated after including publically available full
248
genome sequence data from the additional 11 EHV-1 isolates described above.
249
In SplitsTree4, splits network trees were generated with an uncorrected P characters
250
transformation model, ignoring constant sites. Other models were tested but resulted in no
11
251
significant topological differences in the generated trees. Statistical analyses of the recombination
252
networks were performed using the Phi test (Bruen et al., 2006) as implemented by SplitsTree4. In
253
RDP4 six different methods were used to assess the sequences for recombination breakpoints: RDP,
254
GENECONV, Chimaera, SiScan, MaxChi and Bootscan. Default RDP4 settings were used throughout.
255
Only breakpoints with a Bonferroni-corrected P value <0.05 were reported. Duplicate breakpoints
256
were omitted.
257
258
ACKNOWLEDGEMENTS
259
This work was supported by the Australian Research Council (ARC, DP130103991). JMD is supported
260
by an ARC Future Fellowship (FT140101287
12
261
REFERENCES
262
263
Allen, G., Kydd, J. & Slater, J. (2004). Equid herpesvirus 1 and equid herpesvirus 4 infections. In
264
Infectious diseases of livestock, pp. 829-859. Edited by J. Coetzer & R. Tustin. Newmarket:
265
Oxford University Press.
266
267
268
269
270
271
272
Bagust, T. J. & Pascoe, R. R. (1968). Isolation of equine rhinopnemonitis virus from acute
respiratory disease in a horse in queensland. Aus 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. Aus Vet J 46, 421-427.
Bruen, T. C., Philippe, H. & Bryant, D. (2006). A simple and robust statistical test for detecting the
presence of recombination. Genetics 172, 2665-2681.
Crabb, B. S. & Studdert, M. J. (1990). Comparative studies of the proteins of equine herpesviruses 4
273
and 1 and asinine herpesvirus 3: antibody response of the natural hosts. J Gen Virol 71,
274
2033-2041.
275
276
277
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
278
endometrium, placenta and foetus of equid herpesvirus 1 (EHV-1) abortions. J Comp Pathol
279
104, 379-387.
280
Gilkerson, J. R., Whalley, J. M., Drummer, H. E., Studdert, M. J. & Love, D. N. (1999).
281
Epidemiological studies of equine herpesvirus 1 (EHV-1) in Thoroughbred foals: a review of
282
studies conducted in the Hunter Valley of New South Wales between 1995 and 1997. Vet
283
Microbiol 68, 15-25.
13
284
285
286
287
288
289
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. New Zeal 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.
290
Huson, D. H. (1998). SplitsTree: analyzing and visualizing evolutionary data. Bioinformatics 14, 68-73.
291
Katoh, K. & Standley, D. M. (2013). MAFFT multiple sequence alignment software version 7:
292
improvements in performance and usability. Mol Biol Evol 30, 772-780.
293
Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock, S., Buxton, S., Cooper, A.,
294
Markowitz, S. & other authors (2012). Geneious Basic: an integrated and extendable
295
desktop software platform for the organization and analysis of sequence data.
296
Bioinformatics 28, 1647-1649.
297
298
299
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., Devlin, J. M., Markham, J. F., Noormohammadi, A. H., Browning, G. F., Ficorilli, N. P.,
300
Hartley, C. A. & Markham, P. F. (2013). Phylogenetic and molecular epidemiological studies
301
reveal evidence of multiple past recombination events between infectious laryngotracheitis
302
viruses. PLOS One 8, e55121.
303
Lee, S. W., Markham, P. F., Coppo, M. J., Legione, A. R., Markham, J. F., Noormohammadi, A. H.,
304
Browning, G. F., Ficorilli, N., Hartley, C. A. & Devlin, J. M. (2012). Attenuated vaccines can
305
recombine to form virulent field viruses. Science 337, 188.
14
306
Lee, S. W., Markham, P. F., Markham, J. F., Petermann, I., Noormohammadi, A. H., Browning, G. F.,
307
Ficorilli, N. P., Hartley, C. A. & Devlin, J. M. (2011). First complete genome sequence of
308
infectious laryngotracheitis virus. BMC Genomics 12, 197.
309
Martin, D., Murell, B., Golden, M., Khoosal, A. & Muhire, B. (2015). RDP4: Detection and analysis
310
of recombination patterns in virus genomes. Virus Evolution 1, doi: 10.1093/ve/vev003.
311
Norberg, P., Depledge, D. P., Kundu, S., Atkinson, C., Brown, J., Haque, T., Hussaini, Y., MacMahon,
312
E., Molyneaux, & other authors (2015). Recombination of Globally Circulating Varicella
313
Zoster Virus. J Virol 89, 7133-7146
314
Norberg, P., Liljeqvist, J. A., Bergstrom, T., Sammons, S., Schmid, D. S. & Loparev, V. N. (2006).
315
Complete-genome phylogenetic approach to varicella-zoster virus evolution: genetic
316
divergence and evidence for recombination. J Virol 80, 9569-9576.
317
318
319
320
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.
Peet, R. L., Coackley, W., Smith, V. W. & Main, C. (1978). Equine abortion associated with
herpesvirus. Aus Vet J 54, 151.
321
Peters, G. A., Tyler, S. D., Grose, C., Severini, A., Gray, M. J., Upton, C. & Tipples, G. A. (2006). A
322
full-genome phylogenetic analysis of varicella-zoster virus reveals a novel origin of
323
replication-based genotyping scheme and evidence of recombination between major
324
circulating clades. J Virol 80, 9850-9860.
325
Pignatti, P. F., Cassai, E., Meneguzzi, G., Chenciner, N. & Milanesi, G. (1979). Herpes simplex virus
326
DNA isolation from infected cells with a novel procedure. Virology 93, 260-264.
327
Posada, D. (2002). Evaluation of methods for detecting recombination from DNA sequences:
328
empirical data. Mol Biol Evol 19, 708-717.
15
329
330
331
332
333
334
335
Posada, D. & Crandall, K. A. (2001). Evaluation of methods for detecting recombination from DNA
sequences: computer simulations. PNAS 98, 13757-13762.
Studdert, M. J. (1971). Equine herpesviruses. 4. Concurrent infection in horses with strangles and
conjunctivitis. Aus Vet J 47, 434-436.
Studdert, M. J. & Blackney, M. H. (1979). Equine herpesviruses: on the differentiation of respiratory
from foetal strains of type 1. Aus Vet J 55, 488-492.
Studdert, M. J., Fitzpatrick, D. R., Horner, G. W., Westbury, H. A. & Gleeson, L. J. (1984). Molecular
336
epidemiology and pathogenesis of some equine herpesvirus type 1 (equine abortion virus)
337
and type 4 (equine rhinopneumonitis virus) isolates. Aus Vet J 61, 345-348.
338
Studdert, M. J. & Gleeson, L. J. (1977). Isolation of equine rhinovirus type 1. Aus Vet J 53, 452.
339
Studdert, M. J., Hartley, C. A., Dynon, K., Sandy, J. R., Slocombe, R. F., Charles, J. A., Milne, M. E.,
340
Clarke, A. F. & El-Hage, C. (2003). Outbreak of equine herpesvirus type 1 myeloencephalitis:
341
new insights from virus identification by PCR and the application of an EHV-1-specific
342
antibody detection ELISA. Vet Rec 153, 417-423.
343
Studdert, M. J., Turner, A. J. & Peterson, J. E. (1970). Equine herpesviruses. I. Isolation and
344
characterisation of equine rhinopneumonitis virus and other equine herpesviruses from
345
horses. Aus Vet J 46, 83-89.
346
347
348
349
350
Telford, E. A., Watson, M. S., McBride, K. & Davison, A. J. (1992). The DNA sequence of equine
herpesvirus-1. Virology 189, 304-316.
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, 1197-1203.
Thiry, E., Meurens, F., Muylkens, B., McVoy, M., Gogev, S., Thiry, J., Vanderplasschen, A., Epstein,
351
A., Keil, G. & Schynts, F. (2005). Recombination in alphaherpesviruses. Rev Med Virol 15, 89-
352
103.
16
353
Varrasso, A., Dynon, K., Ficorilli, N., Hartley, C. A., Studdert, M. J. & Drummer, H. E. (2001).
354
Identification of equine herpesviruses 1 and 4 by polymerase chain reaction. Aus Vet J 79,
355
563-569.
356
357
358
Weller, S. K. & Sawitzke, J. A. (2014). Recombination promoted by DNA viruses: phage lambda to
herpes simplex virus. Annu Rev Microbiol 68, 237-258.
Zell, R., Taudien, S., Pfaff, F., Wutzler, P., Platzer, M. & Sauerbrei, A. (2012). Sequencing of 21
359
varicella-zoster virus genomes reveals two novel genotypes and evidence of recombination.
360
J Gen Virol 86, 1608-1622.
361
17
362
363
Table 1. Australian and New Zealand equine herpesvirus isolates used in this study.
Virus ID
Year
Region
Disease association
Site of collection
Reference
GenBank accession
EHV-1
438-77
NZA-77
717A-82
970-90
1029-93
1074-94
1966-02
2019-02
2222-03
3038-07
3045-07
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
Foetus
N/A
Brain or foetus
N/A
Foetus
Foetus
Foetus
Foetus
Foetus
Nasal septum
N/A
(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
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
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
Respiratory tract
Nasal swab
Nasal swab
Nasal swab
Nasal swab
Foetus
Nasal swab
N/A
Foetus
Foetus
Nasal swab
Nasal swab
Nasal swab
Nasal swab
(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
N/A, not available
18
364 Table 2. Recombination breakpoint analysis of EHV-1 and EHV-4 isolates
365
Virus and
region
Breakpoints^
Breakpoint beginning
Breakpoint ending
99% CI
99% CI
Possible viruses involved in
recombination event
Method of breakpoint detection*
EHV-1
IR/TR
116122 - 120856
120865 - 123600
2019-02, 438-77, Unknown
GENECONV, 3Seq, BootScan
EHV-4
UL
UL
UL
IR/TR
US
Undetermined (28)
Undetermined (20689)
Undetermined (96512)
Undetermined (121749)
122044 - 126553
28 – 6954
77657 - 98051
Undetermined (112236)
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, BootScan
MaxChi, Chimaera, SiScan
RDP, MaxChi, BootScan
GENECONV, 3Seq, BootScan,
GENECONV, 3Seq, SiScan, BootScan
366
367
^Nucleotide number according to 438/77 (EHV-1) or 405/76 (EHV-4) reference sequences
368
*Algorithm used to detect recombination, as implemented in RDP4
369
19
370
FIGURE LEGENDS
371
372
Fig. 1. Nucleotide sequence alignment for the complete genomes of Australasian EHV-1 isolates.
373
Alignment of the complete genome sequences of EHV-1 isolates was performed using MAFFT. Our
374
prototype strain 438-77 was used as the reference sequence. Vertical black lines indicate SNPs
375
compared to the reference and dashes indicate sequence gaps.
376
377
Fig. 2. Nucleotide sequence alignment for the complete genomes of Australasian EHV-4 isolates.
378
Alignment of the complete genome sequences of EHV-4 isolates was performed using MAFFT. Our
379
prototype strain EHV-4.405-76 was used as the reference sequence. Vertical black lines indicate
380
SNPs compared to the reference and dashes indicate sequence gaps.
381
382
Fig. 3. Recombination network trees generated from Australasian EHV-4 nucleotide alignments
383
(excluding sequence repeats) using SplitsTree4. (a) Complete genome sequences, (b) unique long
384
region, (c) unique short region and (d) repeat region. The multiple reticulate networks indicate
385
recombination events between the different isolates. The bar indicates the rate of evolution in
386
sequence substitutions per site. The Phi test for detecting recombination, as implemented in
387
SplitsTree4, was highly significant for the whole genome, and for the unique long region, but not for
388
the unique short region or the repeat region.
389
390
Fig. 4. Recombination network trees generated from Australasian EHV-1 nucleotide alignments
391
(excluding sequence repeats) using SplitsTree4. (a) Complete genome sequences, (b) unique long
392
region, (c) unique short region and (d) repeat region. The bar indicates the rate of evolution in
20
393
sequence substitutions per site. The Phi test for detecting recombination, as implemented in
394
SplitsTree4, was not significant for the whole genome, or for any of the individual genome regions.
395
396
397
Fig. S1. Comparison of the genome sequence of our prototype laboratory EHV-1 isolate 438-77 as
398
determined by de novo assembly (438-77 de novo), or by mapping the sequence reads to the EHV-1
399
reference sequence in GenBank accession NC_001491.2 (438-77). Vertical black lines indicate SNPs
400
and dashes indicate sequence gaps. The alignment shows a high level of identity (green) between
401
the sequences generated by the two different methods with the exception of regions of the
402
genome containing repeat sequences (dark blue). These repeat sequences were identified and
403
removed prior to recombination analyses.
404
405
Fig. S2. Comparison of the genome sequence of our prototype laboratory EHV-4 isolate 405-76 as
406
determined by de novo assembly (405-76 de novo) or by mapping the sequence reads to the EHV-4
407
reference sequence in GenBank accession NC_001844.1 (405-76). Vertical black lines indicate SNPs
408
and dashes indicate sequence gaps. The alignment shows a high level of identity (green) between
409
the sequences generated by the two different methods with the exception of regions of the
410
genome containing repeat sequences (dark blue). These repeat sequences were identified and
411
removed prior to recombination analyses.
412
413
Fig. S3. Recombination network trees generated from 11 Australasian and 11 international EHV-1
414
nucleotide alignments (excluding sequence repeats) using SplitsTree4. (a) Complete genome
415
sequences, (b) unique long region, (c) unique short region and (d) repeat region. The multiple
416
reticulate networks indicate recombination events between the different isolates. The bar indicates
21
417
the rate of evolution in sequence substitutions per site. The Phi test for detecting recombination, as
418
implemented in SplitsTree4, was not significant for the whole genome, or for any of the individual
419
genome regions.
420
22