Skip to main content
Erschienen in: Virology Journal 1/2023

Open Access 01.12.2023 | Research

Phylogenomic assessment of 23 equid alphaherpesvirus 1 isolates obtained from USA-based equids

verfasst von: Ugochi Emelogu, Andrew C. Lewin, Udeni B. R. Balasuriya, Chin-Chi Liu, Rebecca P. Wilkes, Jianqiang Zhang, Erinn P. Mills, Renee T. Carter

Erschienen in: Virology Journal | Ausgabe 1/2023

Abstract

Background

Equid alphaherpesvirus 1 (EHV-1) is a global viral pathogen of domestic equids which causes reproductive, respiratory and neurological disease. Few isolates acquired from naturally infected USA-based hosts have been fully sequenced and analyzed to date. An ORF 30 (DNA polymerase) variant (A2254G) has previously been associated with neurological disease in host animals. The purpose of this study was to perform phylogenomic analysis of EHV-1 isolates acquired from USA-based hosts and compare these isolates to previously sequenced global isolates.

Methods

EHV-1 was isolated from 23 naturally infected USA-based equids (6 different states, 15 disease outbreaks) with reproductive (22/23) or neurological disease (1/23). Following virus isolation, EHV-1 DNA was extracted for sequencing using Illumina MiSeq. Following reference-based assembly, whole viral genomes were annotated and assessed. Previously sequenced EHV-1 isolates (n = 114) obtained from global host equids were included in phylogenomic analyses.

Results

The overall average genomic distance was 0.0828% (SE 0.004%) for the 23 newly sequenced USA isolates and 0.0705% (SE 0.003%) when all 137 isolates were included. Clade structure was predominantly based on geographic origin. Numerous nucleotide substitutions (mean [range], 179 [114–297] synonymous and 81 [38–120] non-synonymous substitutions per isolate) were identified throughout the genome of the newly sequenced USA isolates. The previously described ORF 30 A2254G substitution (associated with neurological disease) was found in only one isolate obtained from a host with non-neurological clinical signs (reproductive disease), six additional, unique, non-synonymous ORF 30 substitutions were detected in 22/23 USA isolates. Evidence of recombination was present in most (22/23) of the newly sequenced USA isolates.

Conclusions

Overall, the genomes of the 23 newly sequenced EHV-1 isolates obtained from USA-based hosts were broadly similar to global isolates. The previously described ORF 30 A2254G neurological substitution was infrequently detected in the newly sequenced USA isolates, most of which were obtained from host animals with reproductive disease. Recombination was likely to be partially responsible for genomic diversity in the newly sequenced USA isolates.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s12985-023-02248-z.
Ugochi Emelogu and Andrew C. Lewin have contributed equally to this work.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
A
Adenine
ATCC
American Type Culture Collection
C
Cytosine
CCL
Cellosaurus cell line
CHV-1
Canine herpesvirus type 1
CPE
Cytopathic effect
D
Aspartic acid
DMEM
Dulbecco’s modified eagle medium
DNA
Deoxyribonucleic acid
E
Glutamic acid
EHM
Equine herpesvirus encephalomyelitis
EHV-1
Equine herpesvirus type 1
EHV-8
Equine herpesvirus type 8
F
Phenylalanine
FBS
Fetal bovine serum
FHV-1
Feline herpesvirus type 1
G
Guanine or glycine
K
Lysine
L
Leucine
MAFFT
Multiple alignment using fast fourier transform
MEGA
Molecular evolutionary genetics analysis
N
Asparagine
ORF
Open reading frame
R
Arginine
RDP
Recombination Detection Program
S
Serine
SE
Standard error
T
Thymine
US/USA
United States of America

Background

Equid alphaherpesvirus 1 (EHV-1), a varicellovirus of the Herpesviridae family, is a widespread pathogen of equids which is capable of causing respiratory, ocular, reproductive (i.e., abortion), neurological disease (equine herpes myeloencephalopathy, EHM) and neonatal death [15]. First isolated from abortion material in 1933 [6], EHV-1 is now recognized as a major threat to the global equine industry [7]. In particular, neurological disease in host equids represents a serious welfare and economic problem and has therefore been the focus of numerous prior investigations [811].
EHV-1 is a double stranded DNA virus with an approximately 150 kilobase genome consisting of 76 unique open reading frames (ORF) and 4 duplicated repeat regions [12]. Previous studies have indicated that ORF 30 (DNA polymerase) substitution A2254G (leading to amino acid variation N752D) is associated with neurological disease in host animals [9, 1315]. Although the underlying mechanism for this is unclear, it has been suggested that EHV-1 with the A2254G substitution replicates at a higher level and induces a longer-lasting viraemia than isolates without this substitution [16]. Notably, isolates recovered from host animals with neurological disease do not consistently possess this substitution [17]. Although prior work has found that certain isolates are more capable of inducing abortion in experimental studies [18, 19], EHV-1 genetic determinants for abortion have not been identified. The ORF 68 region (which encodes a non-essential membrane associated component) has previously been utilized as a genetic marker for grouping EHV-1 isolates [20].
Although several phylogenomic assessments of naturally occurring EHV-1 have been previously performed, limited numbers of viral isolates obtained from USA-based host animals have been included in these analyses [17, 21, 22]. Ongoing phylogenomic investigation of existing isolates is likely to improve global surveillance and subsequent control of this virus. The purpose of this study was to perform phylogenomic analysis of EHV-1 isolates acquired from USA-based hosts, and to compare these isolates to previously sequenced global isolates. Based on previous work, we hypothesized that isolates would form clades based on geographic origin [17] and that the previously described ORF 30 A2254G substitution would be present only in isolates obtained from host animals with neurological disease [20].

Methods

Viral isolates and hosts

Twenty-three archived EHV-1 isolates obtained over a 24-year period (1997–2021) from naturally infected, USA-based domestic equids were sequenced in this study. Isolates from host animals located in 6 different US states (i.e., California, Iowa, Indiana, Virginia, North Dakota and South Dakota) were included. The viral isolates sequenced in this study (with associated host information, where available) are shown in Table 1.
Table 1
Viral isolate and associated host data for the 23 EHV-1 isolates sequenced (whole viral genome) for the present study
Isolate
Genbank Accession ID
Host State
Collection Date
Isolation Date
Tissue source
Host Age
Host Species
Host Sex
Disease Type
ORF30 A > G, 2254
Synonymous Substitutions
Non-Synonymous Substitutions
LS0433182
OR085515
South Dakota
Unknown
9/1/2004
Unknown
8 years
Equus caballus
Female
Neurological
No
138
62
LS050627
OR085496
North Dakota
Unknown
6/27/2005
Unknown
Unknown
Unknown
Unknown
Reproductive
Yes
135
71
LS080124
OR085497
Indiana
Unknown
1/24/2008
Unknown
Unknown
Unknown
Unknown
Reproductive
No
174
86
LS100106
OR085498
Indiana
Unknown
1/6/2010
Unknown
Unknown
Unknown
Unknown
Reproductive
No
174
86
LS110097
OR085499
Iowa
3/25/2011
11/7/2021
Liver
Fetal
Unknown
Unknown
Reproductive
No
174
87
LS110114
OR085500
Iowa
4/6/2011
11/7/2021
Lung
Fetal
Unknown
Unknown
Reproductive
No
115
58
LS110976
OR085501
Iowa
3/25/2011
11/7/2021
Spleen
Fetal
Unknown
Unknown
Reproductive
No
174
87
LS111457
OR085502
Iowa
4/6/2011
11/7/2021
Liver
Fetal
Unknown
Unknown
Reproductive
No
132
43
LS113812
OR085503
Iowa
2/4/2011
11/7/2021
Liver
Fetal
Unknown
Unknown
Reproductive
No
157
83
LS119764
OR085504
Iowa
3/25/2011
11/7/2021
Lung
Fetal
Unknown
Unknown
Reproductive
No
116
49
LS130922
OR085505
Iowa
1/8/2013
3/4/2013
Unknown
Fetal
Unknown
Unknown
Reproductive
No
174
71
LS140310
OR085506
Indiana
Unknown
3/10/2014
Unknown
Unknown
Unknown
Unknown
Reproductive
No
236
111
LS140325
OR085507
Indiana
Unknown
3/25/2014
Unknown
Unknown
Unknown
Unknown
Reproductive
No
158
85
LS143101
OR085508
Indiana
Unknown
3/10/2014
Unknown
Unknown
Unknown
Unknown
Reproductive
No
228
115
LS161221
OR085509
Indiana
Unknown
12/21/2016
Unknown
Unknown
Unknown
Unknown
Reproductive
No
191
88
LS161227
OR085510
Indiana
Unknown
12/27/2016
Unknown
Unknown
Unknown
Unknown
Reproductive
No
227
91
LS170109
OR085511
Indiana
Unknown
1/9/2017
Unknown
Unknown
Unknown
Unknown
Reproductive
No
261
104
LS170307
OR085512
Indiana
Unknown
3/7/2017
Unknown
Unknown
Unknown
Unknown
Reproductive
No
145
93
LS180416
OR085513
Indiana
Unknown
4/16/2018
Unknown
Unknown
Unknown
Unknown
Reproductive
No
297
120
LS213627
OR085514
Iowa
4/22/2021
6/8/2021
Lung
Unknown
Unknown
Unknown
Reproductive
No
126
52
LS9719959
OR085516
Virginia
Unknown
4/15/1997
Unknown
Fetal
Equus caballus
Unknown
Reproductive
No
114
38
LS9816616
OR085517
California
Unknown
3/16/1998
Unknown
Fetal
Equus caballus
Male
Reproductive
No
294
112
LS9922376
OR085518
Virginia
Unknown
5/5/1999
Unknown
Fetal
Equus caballus
Female
Reproductive
No
184
77
Note that complete host data was not available in most cases. ORF30 A > G, 2254 refers to a previously described substitution which has been associated with neurological disease [20]
A total of 114 additional EHV-1 isolate genomes which had previously been fully/near fully sequenced were obtained from Genbank (https://​www.​ncbi.​nlm.​nih.​gov/​genbank/​) for inclusion in phylogenomic analyses. These isolates were obtained from hosts located in four global regions including Europe (84/114), USA (6/114), Australia (11/114) and Asia (13/114). A list of these isolates can be found in Additional file 1: Table S1.

Cell culture and viral isolation

CCL-57 cells (Equine dermal cells, ATCC CCL-57, Manassas, VA) were cultured at 37 °C/5% CO2 in T25 flasks (Thermo Fisher, Waltham, MA, USA) using Dulbecco’s modified Eagle medium (DMEM) (Thermo Fisher, Waltham, MA, USA) fortified with 10% fetal bovine serum (FBS) (Thermo Fisher, Waltham, MA, USA) and 1% penicillin/streptomycin (Thermo Fisher, Waltham, MA, USA) until 90–100% confluent.
Approximately 100µL of each viral stock (virus previously isolated using various cell lines for archival) was added to 1 mL of DMEM fortified with 2% FBS and 1% penicillin/streptomycin sulfate and briefly vortexed before being added to individual T25 flasks containing confluent CCL-57 cells. Flasks were then placed on a rocker at room temperature for 60 min, after which an additional 4 mL of DMEM fortified with 2% FBS and 1% penicillin/streptomycin sulfate was added to each flask before incubation at 37 °C and 5% CO2. Flasks were then monitored daily for visible cytopathic effect (CPE). Once 100% CPE was verified, three cycles of freezing at − 80 °C and thawing at room temperature were carried out. The contents of each flask were transferred to a 15 ml conical tube and centrifuged at 400 × g for 5 min at 4 °C. The resultant supernatant was then transferred to cryotubes and immediately stored at − 80 °C pending DNA extraction.

Viral DNA extraction, viral species confirmation and DNA concentration assessment

Using 200µL of stored cell culture supernatant, viral DNA extraction was carried out using a commercially available kit (Purelink Viral DNA Mini Kit, Thermo Fisher, Waltham, MA, USA). DNA concentration was determined using a Qubit dsDNA HS Assay Kit (Life Technologies, Grand Island, NY). Prior to full viral genome sequencing, real-time polymerase chain reaction (7900HT Fast Real-Time PCR System, Applied Biosystems) was carried out on samples to confirm viral identity, using EHV-1 specific primers and probe [23].

Illumina genomic DNA sequencing

Exact DNA concentrations for all samples were used to prepare sequencing libraries using the Nextera XT DNA Library Prep Kit (Illumina Inc. San Diego, CA, USA) and Nextera XT Index Kit v2 (Illumina Inc. San Diego, CA, USA). Final library quality and quantity were determined using a Fragment Analyzer Instrument (Fragment Analyzer System, Agilent). Libraries were pooled following indexing, prior to sequencing. Paired-end whole genome sequencing was performed using an Illumina MiSeq instrument using a 600 cycles MiSeq Reagent Kit v3 (Illumina Inc. San Diego, CA, USA).

Genome assembly and alignments

Reference-based genome (V592, Genbank accession number of AY464052) assembly was carried out using Geneious Prime (version 2020.2.4), as previously described [24, 25]. In brief, paired end reads were first trimmed using BBDuk Adapter/Quality trimmer version 38.84 (right end, Kmer length = 27, maximum substitution = 1, minimum quality = 30, minimum overlap = 20, minimum length = 30). The trimmed paired end reads were then mapped to the reference genome using Geneious Prime. A consensus sequence was generated from the aligned reads with gaps filled with “N’s”. Each genome was annotated using annotation similarity transfer within Geneious Prime, prior to submission to the online Genbank data repository.
Viral genomes were aligned as previously described [24, 25], using MAFFT alignment tool (MAFFT ver 7.490), with default parameters [26]. Multiple alignments (with or without an equid alphaherpesvirus 8 (EHV-8) outgroup, Genbank accession NC017826) were created to include all of the newly sequenced USA isolates with or without previously sequenced isolates (obtained from Genbank, shown in Additional file 1: Table S1). Sites with at least 20% gaps were stripped from the alignments using the ‘Mask Alignment’ tool in Geneious Prime for subsequent phylogenomic analyses.

Nuclotide subsitution analysis

Nucleotide substitution analysis was performed as previously described [24, 25] using the ‘Geneious Variant Finder’ (Geneious Prime version 2020.2.4, minimum coverage = 100, minimum variant frequency = 0.25, maximum variant p value = 10–6). Substitutions were identified by comparing the EHV-1 sequenced isolates to the reference genome (V592).

Phylogenomic and recombination analysis

Phylogenomic analysis of whole viral genomes was performed as previously described [24, 25]. ModelFinder [27], within IQ-Tree 2 version 1.6.12 [28], was used for automatic selection of the best-fit model (K3Pu + F + R9) for the stripped alignment containing all available EHV-1 isolates and an EHV-8 outgroup (NC017826). The resultant treefile was viewed using Splitstree (version 4.16.1) [29] and Geneious Prime. Pairwise genomic distances were determined using EHV-1 alignments (without an EHV-8 outgroup) in MEGA11 (ver. 11.0.13) [30] with the gamma distribution model, partial deletion of gaps and 1000 bootstrap replicates.
Recombination analysis was performed using RDP version 4.100 [31] on an alignment containing all 137 EHV-1 genomes (without an EHV-8 outgroup) using manual bootscan (window = 1200, step = 500, replicates = 100, 70% cutoff, Jin and Nei model [32]), RDP [33], GENECONV [34], MaxChI [35], Chimaera [36] and Siscan [37].

Results

Viral Isolates, host information and sequencing results

The whole genome of 23 EHV-1 isolates from 23 host animals in six US states were sequenced; 10/23 from Indiana, 8/23 from Iowa, 2/23 from Virginia, 1 each from California, North Dakota, and South Dakota (Table 1). Isolates were collected and archived over a period of 24 years between 1997 and 2021. In total, isolates from 15 separate outbreaks (same geographic location, isolated or collected within 30 days of other isolate(s)) were included. Host data was included, where available. The age of 11/23 hosts and species (Equus caballus) of 4/23 hosts was known, while sex of 3/23 hosts was known (Table 1). Host disease data was available for all 23 isolates (Table 1). Of the 23 host animals, 22/23 had reproductive disease (including abortion and stillbirth) while 1/23 had neurological disease (ataxia and urinary incontinence).
Additional file 2: Table S2 shows sequencing details for the 23 newly sequenced EHV-1 isolates. The total number of reads obtained ranged from 340,648 (LS9922376) to 1,946,622 (LS140310). The average number of mapped reads ranged from 69,798 for strain LS110976 to 452,971 for strain LS143101. Mean genome coverage was 240X, with a range of 62.1X (LS110097) to 607.9X (LS143101). The overall mean genome GC content was 55.3%, with a range of 54.3% (LS050627) to 66.5% (LS110097). The average mapped genome length was 149,920 with a range of 148,164 (LS110097) to 151,048 (LS180416).

Phylogenomic analysis

The phylogenomic relationships between all 137 (114 previously sequenced, 23 newly sequenced) sequenced EHV-1 isolates (with an EHV-8 outgroup) are shown in Fig. 1. A maximum likelihood tree, generated using the same treefile, is shown in Fig. 2. Newly sequenced USA based isolates were widely distributed among existing clades, with some grouping observed between certain isolates obtained from the same disease outbreak. For example, 10 newly sequenced isolates from Indiana were analyzed, obtained from 6 separate disease outbreaks. Of these 6 outbreaks, 2 outbreaks yielded 3 isolates each; one in 2014 (LS140310, LS140325 and LS143101) and the other in 2016/2017 (LS161221, LS161227 and LS170109). Isolates from each outbreak were found to be mostly identical, as expected, and clustered closely together. Of the 6 isolates from a single outbreak in Iowa in 2011, 5/6 of these were mostly homologous and clustered together (LS110097, LS110114, LS110976, LS111457 and LS119764), with the remaining isolate (LS113812) found at a distant position (see ‘Distance Analysis’, below). Although many of the 137 isolates grouped according to geographic origin (Europe, USA, Australia or Asia), numerous exceptions were observed (Figs. 1 and 2).

Distance analysis

The overall average genomic distance was 0.0828% (SE 0.004%) for the 23 newly sequenced USA isolates and 0.0705% (SE 0.003%) when all 137 isolates were included. Interregional genomic distances are shown in Table 2. Overall, the interregional distances as shown in Table 2 were broadly similar. The lowest distance difference was between isolates from Australia and Europe (0.0592%) and the greatest distance difference was between Asia and United States (0.0913%). Intraregional genomic distances were; 0.0828% (SE 0.004%) for USA isolates, 0.0784% (SE 0.004%) for Asian isolates, 0.0580% (SE 0.003%) for European isolates and 0.0574% (SE 0.004%) for Australian isolates.
Table 2
Mean interregional genomic distances of isolates obtained in Europe, USA, Asia or Australia
 
Europe
USA
Asia
Australia
Europe
 
0.00392
0.00407
0.00341
USA
0.0813
 
0.00425
0.00399
Asia
0.0816
0.0913
 
0.00426
Australia
0.0592
0.0750
0.0803
 
The lower left values are the interregional distance values (expressed as percentages) and the upper right values are the corresponding standard error values
Of the six Iowa isolates (LS110097, LS111457, LS113812, LS110976, LS119764 and LS110114) that originated from the same outbreak, five isolates clustered together (LS110097, LS111457, LS110976, LS119764 and LS110114) with a relatively low intraclade distance of 0.00763% (SE 0.00173%). When the remaining isolate from the same outbreak was considered (LS113812), inter-isolate distance was approximately 5 × greater; 0.0367% (SE 0.00389%).

Nucleotide substitution detection

Nucleotide substitutions (relative to the reference genome, V592) were detected in all 23 newly sequenced EHV-1 isolates. On average, 179 (range 114–297) synonymous and 81 (range 38–120) non-synonymous substitutions were identified in each genome (Table 1). Nucleotide substitutions were relatively evenly distributed throughout the EHV-1 genomes. Details of substitutions in each of the 23 newly sequenced EHV-1 isolates can be found in Additional file 3: Tables S3 to S25.
Following whole genome substitution detection in each of the 23 newly sequenced EHV-1 isolates, genomes were specifically assessed for the previously described ORF 30 A2254G substitution, which has been associated with host neurological disease [20] (Table 1). This substitution was present in only 1/23 newly sequenced isolates (LS050627), which was obtained from a host equid with reproductive disease located in North Dakota in 2005. The genome of the single newly-sequenced isolate (LS0433182) from a host animal with confirmed neurological disease did not contain the ORF 30 A2254G substitution. In addition to the ORF 30 A2254G substitution, a total of 6 additional, unique non-synonymous ORF 30 substitutions were identified in the 23 newly sequenced USA isolates (Table 3). Most (22/23) isolates contained at least 1 (range 0–3) of these 6 unique non-synonymous substitutions. These 6 unique non-synonymous ORF 30 substitutions included; A2279G (D760G), A2968G (K990E), A1984C (S662R), G1286A (R429K), T2149C (F717L) and C3131T (S1044L). Five unique synonymous ORF 30 substitutions were identified in the 23 newly sequences USA isolates which included; T924C, G96A, G2805T, G2874A and C2352T.
Table 3
Synonymous and non-synonymous ORF30 (DNA polymerase) substitutions in each of the 23 newly sequenced USA EHV-1 isolates
Isolate
ORF 30 synonymous substitutions
ORF 30 non-synonymous substitutions
ORF location
Polymorphism
ORF location
Polymorphism
AA change, position
LS0433182
924
T > C
2279
A > G
D to G, 760
   
2968
A > G
K to E, 990
LS050627
924
T > C
2254
A > G
N to D, 752
   
2968
A > G
K to E, 990
LS080124
924
T > C
1984
A > C
S to R, 662
   
2968
A > G
K to E, 990
LS100106
924
T > C
2968
A > G
K to E, 990
LS110097
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2805
G > T
   
LS110114
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2805
G > T
   
LS110976
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2805
G > T
   
LS111457
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2805
G > T
   
LS113812
924
T > C
2968
A > G
K to E, 990
LS119764
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2805
G > T
   
LS130922
96
G > A
2968
A > G
K to E, 990
 
924
T > C
   
 
2874
G > A
   
LS140310
924
T > C
1286
G > A
R to K, 429
   
2968
A > G
K to E, 990
LS140325
924
T > C
2968
A > G
K to E, 990
LS143101
924
T > C
1286
G > A
R to K, 429
   
2968
A > G
K to E, 990
LS161221
924
T > C
2968
A > G
K to E, 990
LS161227
924
T > C
2968
A > G
K to E, 990
LS170109
924
T > C
2968
A > G
K to E, 990
LS170307
924
T > C
1984
A > C
S to R, 662
   
2968
A > G
K to E, 990
LS180416
924
T > C
2968
A > G
K to E, 990
 
2352
C > T
   
LS213627
924
T > C
2968
A > G
K to E, 990
LS9719959
96
G > A
N/A
  
 
924
T > C
   
LS9816616
924
T > C
2149
T > C
F to L, 717
   
2968
A > G
K to E, 990
   
3131
C > T
S to L, 1044
LS9922376
924
T > C
2968
A > G
K to E, 990
Polymorphisms: A = adenine, G = guanine, C = cytosine, T = thymine. Amino acid abbreviations: D = aspartic acid, G = glycine, K = lysine, E = glutamic acid, N = asparagine, S = serine, R = arginine, F = phenylalanine, L = leucine

Recombination analysis

Genomic evidence of recombination was present in most (22/23) of the 23 newly sequenced EHV-1 isolates, when assessed using RDP4 [33]. Manual bootscan assessment detected evidence of recombination involving at least 1 other EHV-1 isolate in 20/23 newly sequenced isolates (Additional file 4: Table S26). The mean number of possible recombinants per isolate was 16 (range 0–53) based on manual bootscan assessment. When the complete RDP analysis was utilized, most (16/23) isolates were found to demonstrate evidence of recombination by at least one of the 5 additional methods of detection (RDP, GENECONV, MaxChi, Chimaera, Siscan); details can be found in Additional file 5: Table S27. Only 1 isolate (LS050627) was found to demonstrate no evidence of recombination by any of the methods utilized. As noted above (see Nucleotide Substitution Detection), this isolate was obtained in 2005 from an equid host located in North Dakota with reproductive disease and possessed the previously described ORF 30 A2254G neurological substitution.

Discussion

Phylogenomic and recombinational assessment of the 23 newly sequenced EHV-1 isolates obtained from USA-based hosts revealed many similarities with previously sequenced global isolates. Phylogenomic assessment of previously sequenced EHV-1 isolates and related alphaherpesviruses have been reported, facilitating comparison with the results presented herein.
As previously described [17], the genome of EHV-1 isolates obtained from multiple hosts involved in the same disease outbreak were found to share a high degree of similarity. This finding is not unexpected and has previously been reported in studies of a similar nature for related alphaherpesviruses such as canid alphaherpesvirus 1 (CHV-1) [24] and felid alphaherpesvirus 1 (FHV-1) [25]. Within our sample set, multiple isolates from three notable outbreaks of reproductive disease were included: Iowa in 2011, Indiana in 2014 and 2016. In all 3 cases, most isolates from the same outbreak (with the exception of LS113812, Iowa 2011) were found to be near-identical. While isolates obtained from hosts in the same geographic region (USA, Asia, Australia or Europe) clustered together, there were numerous exceptions to this pattern which was observed. For example, newly sequenced USA isolates clustered with varying combinations of isolates from distant geographic locations, including Europe, Australia and Asia. While small variations in pairwise interregional genomic distances were detected, it is clear that EHV-1 isolate geographic origin cannot be determined solely on sequence data. In contrast, FHV-1 isolates obtained from widespread geographic locations mostly do form clades based on geographic origin [38]. Although the definitive reason for this discrepancy is unknown, possible reasons include sample size, degree of viral intraspecies conservation and host species differences in global animal (or animal product) movements. Although previous assessments [20, 39, 40] have sought to utilize EHV-1 ORF 68 as the primary method to classify isolates into clade structures, we chose to perform viral full genome sequencing to identify nucleotide substitutions throughout the genome, as has been previously performed [17] and described.
Since being described almost 20 years ago [20], the ORF 30 A2254G substitution has been investigated in the context of equine herpes myeloencephalopathy (EHM) [2, 9, 1315, 17, 41, 42]. Although the substitution has been shown to be significantly associated with neurological disease [20], notably the substitution is not present in all isolates recovered from hosts with EHM [17]. As the underlying mechanism for this association between this specific substitution and EHM is unknown, the reasons for this apparent inconsistency are presently unknown. Most (22/23) of the newly sequenced USA isolates included in the present study were obtained from host animals with reproductive disease, with only 1 originating from a host equid with neurological disease. As noted above, only 1/23 isolates possessed the ORF 30 A2254G substitution, which originated from a host with reproductive disease. The single isolate obtained from the host animal with neurological disease did not possess the ORF 30 A2254G substitution. Inclusion of a higher number of isolates originating from hosts with neurological disease in future studies of similar design are suggested.
Several other ORF 30 substitutions have been described, in addition to the well-known ORF 30 A2254G substitution. Certain EHV-1 isolates (obtained from animals with reproductive disease) have been noted to contain a non-synonymous substitution in ORF 30 at position 2258 (A2258C) [43]; this substitution was not detected in any of the isolates in our sample. In addition, a recently described non-synonymous ORF 30 substitution at position 2254 (A2254C) has been described and documented in isolates from the USA and France [4446]; again, this substitution was not detected in any of the isolates in our sample. In addition to the ORF 30 A2254G substitution, we detected 6 other unique non-synonymous substitutions in the 23 newly sequenced isolates, some of which have been previously described [17, 20]. Surveillance using assays targeting this highly conserved region is likely to be beneficial for use in future outbreaks. The effect of each of these substitutions on virulence, if any, is unknown.
Both intraspecies and interspecies recombination have been shown to be prevalent mechanisms of diversity in EHV-1 and related alphaherpesviruses [17, 24, 25, 38, 4751]. We detected evidence of recombination in most (22/23) of the newly sequenced EHV-1 isolates in the present study. It is unknown why one of the newly sequenced isolates (LS050627) did not demonstrate evidence of recombination, although this could represent the limitations of the predictive computational processes utilized for this purpose. Increased understanding of the mechanisms by which EHV-1 substitutions develop is likely to have significant implications for both disease surveillance and control in host equids.
The present study has several limitations, including sample size, incomplete host animal information, lack of host disease type diversity, and viral genome gaps following sequencing. The present study included 23 newly sequenced EHV-1 isolates from USA based host animals. While concerted attempts were made to include a higher number of isolates, sample availability was a limiting factor. Approximately 46 samples (suspected to contain either EHV-1 DNA or viable EHV-1) were initially screened, but only 23 yielded high quality EHV-1 DNA in quantities suitable for Illumina MiSeq sequencing. During sample collection, efforts were made to collect as much host animal data as possible. As many of the samples had been collected and archived many years prior, host information was not consistently readily available. Most (22/23) of the EHV-1 isolates were obtained from hosts with reproductive disease. Inclusion of a higher number of isolates from animals with neurological and/or respiratory disease may have facilitated assessment of relationships between viral genome substitutions and host disease type. Finally, in common with all Illumina platform viral genome sequencing assessments, the reference-based assembly of high GC content regions resulted in sequence gaps with the genome sequences of the isolates from this study. While unavoidable with this approach, this was accounted for during analysis and is therefore not expected to have affected the results or subsequent conclusions.

Conclusions

Overall, the genomes of the 23 newly sequenced EHV-1 isolates obtained from USA-based hosts were similar to previously sequenced global isolates. The previously described ORF30 A2254G substitution was infrequently detected in the newly sequenced USA isolates, most of which were obtained from host animals with reproductive disease. In line with previous findings, recombination was likely to be partially responsible for genomic diversity in the newly sequenced USA isolates.

Acknowledgements

The authors would like to thank the following persons for assistance with sample acquisition and storage: Nikole Ineck (LSU), Hannah Gafen (LSU), Melanie Mironovich (LSU), Jeremy Snyder (NJ Department of Agriculture), Mary Lea Killian (NVSL), Mia Torchetti (NVSL), The Infectious Agent Repository at the Minnesota Veterinary Diagnostic Laboratory and The Diagnostic Virology Laboratory at the National Veterinary Services Laboratory. In addition, the authors would like to thank Vladimir Chouljenko (LSU) for assistance with viral genome sequencing.

Declarations

All data was generated using previously stored samples and existing sequencing data.
Not applicable.

Competing interests

The authors declare that they have no competing interests.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Damiani AM, de Vries M, Reimers G, Winkler S, Osterrieder N. A severe equine herpesvirus type 1 (EHV-1) abortion outbreak caused by a neuropathogenic strain at a breeding farm in northern Germany. Vet Microbiol. 2014;172(3–4):555–62.PubMedCrossRef Damiani AM, de Vries M, Reimers G, Winkler S, Osterrieder N. A severe equine herpesvirus type 1 (EHV-1) abortion outbreak caused by a neuropathogenic strain at a breeding farm in northern Germany. Vet Microbiol. 2014;172(3–4):555–62.PubMedCrossRef
2.
Zurück zum Zitat Negussie H, Gizaw D, Tessema TS, Nauwynck HJ. Equine herpesvirus-1 myeloencephalopathy, an emerging threat of working equids in Ethiopia. Transbound Emerg Dis. 2017;64(2):389–97.PubMedCrossRef Negussie H, Gizaw D, Tessema TS, Nauwynck HJ. Equine herpesvirus-1 myeloencephalopathy, an emerging threat of working equids in Ethiopia. Transbound Emerg Dis. 2017;64(2):389–97.PubMedCrossRef
3.
Zurück zum Zitat Schulman ML, Becker A, van der Merwe BD, Guthrie AJ, Stout TA. Epidemiology and reproductive outcomes of EHV-1 abortion epizootics in unvaccinated Thoroughbred mares in South Africa. Equine Vet J. 2015;47(2):155–9.PubMedCrossRef Schulman ML, Becker A, van der Merwe BD, Guthrie AJ, Stout TA. Epidemiology and reproductive outcomes of EHV-1 abortion epizootics in unvaccinated Thoroughbred mares in South Africa. Equine Vet J. 2015;47(2):155–9.PubMedCrossRef
4.
Zurück zum Zitat Hussey GS, Goehring LS, Lunn DP, Hussey SB, Huang T, Osterrieder N, et al. Experimental infection with equine herpesvirus type 1 (EHV-1) induces chorioretinal lesions. Vet Res. 2013;44(1):118.PubMedPubMedCentralCrossRef Hussey GS, Goehring LS, Lunn DP, Hussey SB, Huang T, Osterrieder N, et al. Experimental infection with equine herpesvirus type 1 (EHV-1) induces chorioretinal lesions. Vet Res. 2013;44(1):118.PubMedPubMedCentralCrossRef
5.
Zurück zum Zitat Vandekerckhove AP, Glorieux S, Gryspeerdt AC, Steukers L, Van Doorsselaere J, Osterrieder N, et al. Equine alphaherpesviruses (EHV-1 and EHV-4) differ in their efficiency to infect mononuclear cells during early steps of infection in nasal mucosal explants. Vet Microbiol. 2011;152(1–2):21–8.PubMedCrossRef Vandekerckhove AP, Glorieux S, Gryspeerdt AC, Steukers L, Van Doorsselaere J, Osterrieder N, et al. Equine alphaherpesviruses (EHV-1 and EHV-4) differ in their efficiency to infect mononuclear cells during early steps of infection in nasal mucosal explants. Vet Microbiol. 2011;152(1–2):21–8.PubMedCrossRef
6.
Zurück zum Zitat Dimock WW, Edwards PR. Is there a filterable virus of abortion in mares? Ky Agric Exp Stn Bull. 1933;333(9):297–301. Dimock WW, Edwards PR. Is there a filterable virus of abortion in mares? Ky Agric Exp Stn Bull. 1933;333(9):297–301.
7.
Zurück zum Zitat Ma G, Azab W, Osterrieder N. Equine herpesviruses type 1 (EHV-1) and 4 (EHV-4)-masters of co-evolution and a constant threat to equids and beyond. Vet Microbiol. 2013;167(1–2):123–34.PubMedCrossRef Ma G, Azab W, Osterrieder N. Equine herpesviruses type 1 (EHV-1) and 4 (EHV-4)-masters of co-evolution and a constant threat to equids and beyond. Vet Microbiol. 2013;167(1–2):123–34.PubMedCrossRef
8.
Zurück zum Zitat Fritsche AK, Borchers K. Detection of neuropathogenic strains of Equid Herpesvirus 1 (EHV-1) associated with abortions in Germany. Vet Microbiol. 2011;147(1–2):176–80.PubMedCrossRef Fritsche AK, Borchers K. Detection of neuropathogenic strains of Equid Herpesvirus 1 (EHV-1) associated with abortions in Germany. Vet Microbiol. 2011;147(1–2):176–80.PubMedCrossRef
9.
Zurück zum Zitat Pronost S, Léon A, Legrand L, Fortier C, Miszczak F, Freymuth F, et al. Neuropathogenic and non-neuropathogenic variants of equine herpesvirus 1 in France. Vet Microbiol. 2010;145(3–4):329–33.PubMedCrossRef Pronost S, Léon A, Legrand L, Fortier C, Miszczak F, Freymuth F, et al. Neuropathogenic and non-neuropathogenic variants of equine herpesvirus 1 in France. Vet Microbiol. 2010;145(3–4):329–33.PubMedCrossRef
10.
Zurück zum Zitat Stasiak K, Rola J, Ploszay G, Socha W, Zmudzinski JF. Detection of the neuropathogenic variant of equine herpesvirus 1 associated with abortions in mares in Poland. BMC Vet Res. 2015;11:102.PubMedPubMedCentralCrossRef Stasiak K, Rola J, Ploszay G, Socha W, Zmudzinski JF. Detection of the neuropathogenic variant of equine herpesvirus 1 associated with abortions in mares in Poland. BMC Vet Res. 2015;11:102.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Tsujimura K, Oyama T, Katayama Y, Muranaka M, Bannai H, Nemoto M, et al. Prevalence of equine herpesvirus type 1 strains of neuropathogenic genotype in a major breeding area of Japan. J Vet Med Sci. 2011;73(12):1663–7.PubMedCrossRef Tsujimura K, Oyama T, Katayama Y, Muranaka M, Bannai H, Nemoto M, et al. Prevalence of equine herpesvirus type 1 strains of neuropathogenic genotype in a major breeding area of Japan. J Vet Med Sci. 2011;73(12):1663–7.PubMedCrossRef
12.
Zurück zum Zitat Telford EA, Watson MS, McBride K, Davison AJ. The DNA sequence of equine herpesvirus-1. Virology. 1992;189(1):304–16.PubMedCrossRef Telford EA, Watson MS, McBride K, Davison AJ. The DNA sequence of equine herpesvirus-1. Virology. 1992;189(1):304–16.PubMedCrossRef
13.
Zurück zum Zitat Castro ER, Arbiza J. Detection and genotyping of equid herpesvirus 1 in Uruguay. Rev Sci Tech. 2017;36(3):799–806.PubMedCrossRef Castro ER, Arbiza J. Detection and genotyping of equid herpesvirus 1 in Uruguay. Rev Sci Tech. 2017;36(3):799–806.PubMedCrossRef
14.
Zurück zum Zitat Goodman LB, Loregian A, Perkins GA, Nugent J, Buckles EL, Mercorelli B, et al. A point mutation in a herpesvirus polymerase determines neuropathogenicity. PLoS Pathog. 2007;3(11): e160.PubMedPubMedCentralCrossRef Goodman LB, Loregian A, Perkins GA, Nugent J, Buckles EL, Mercorelli B, et al. A point mutation in a herpesvirus polymerase determines neuropathogenicity. PLoS Pathog. 2007;3(11): e160.PubMedPubMedCentralCrossRef
15.
Zurück zum Zitat Khusro A, Aarti C, Rivas-Caceres RR, Barbabosa-Pliego A. Equine herpesvirus-I infection in horses: recent updates on its pathogenicity, vaccination, and preventive management strategies. J Equine Vet Sci. 2020;87: 102923.PubMedCrossRef Khusro A, Aarti C, Rivas-Caceres RR, Barbabosa-Pliego A. Equine herpesvirus-I infection in horses: recent updates on its pathogenicity, vaccination, and preventive management strategies. J Equine Vet Sci. 2020;87: 102923.PubMedCrossRef
16.
Zurück zum Zitat Allen GP, Breathnach CC. Quantification by real-time PCR of the magnitude and duration of leucocyte-associated viraemia in horses infected with neuropathogenic vs. non-neuropathogenic strains of EHV-1. Equine Vet J. 2006;38(3):252–7.PubMedCrossRef Allen GP, Breathnach CC. Quantification by real-time PCR of the magnitude and duration of leucocyte-associated viraemia in horses infected with neuropathogenic vs. non-neuropathogenic strains of EHV-1. Equine Vet J. 2006;38(3):252–7.PubMedCrossRef
17.
Zurück zum Zitat Bryant NA, Wilkie GS, Russell CA, Compston L, Grafham D, Clissold L, et al. Genetic diversity of equine herpesvirus 1 isolated from neurological, abortigenic and respiratory disease outbreaks. Transbound Emerg Dis. 2018;65(3):817–32.PubMedPubMedCentralCrossRef Bryant NA, Wilkie GS, Russell CA, Compston L, Grafham D, Clissold L, et al. Genetic diversity of equine herpesvirus 1 isolated from neurological, abortigenic and respiratory disease outbreaks. Transbound Emerg Dis. 2018;65(3):817–32.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Mumford J, Hannant D, Jesset D, O’Neill T, Smith K, Ostlund E. Abortigenic and neurological disease caused by experimental infection with equid herpesvirus-1. In: Nakajima H, Plowright W, editors. International conference on equine infectious diseases: Newmarket. R&W Publications; 1994. p. 261–75. Mumford J, Hannant D, Jesset D, O’Neill T, Smith K, Ostlund E. Abortigenic and neurological disease caused by experimental infection with equid herpesvirus-1. In: Nakajima H, Plowright W, editors. International conference on equine infectious diseases: Newmarket. R&W Publications; 1994. p. 261–75.
19.
Zurück zum Zitat Gardiner DW, Lunn DP, Goehring LS, Chiang YW, Cook C, Osterrieder N, et al. Strain impact on equine herpesvirus type 1 (EHV-1) abortion models: viral loads in fetal and placental tissues and foals. Vaccine. 2012;30(46):6564–72.PubMedCrossRef Gardiner DW, Lunn DP, Goehring LS, Chiang YW, Cook C, Osterrieder N, et al. Strain impact on equine herpesvirus type 1 (EHV-1) abortion models: viral loads in fetal and placental tissues and foals. Vaccine. 2012;30(46):6564–72.PubMedCrossRef
20.
Zurück zum Zitat Nugent J, Birch-Machin I, Smith KC, Mumford JA, Swann Z, Newton JR, et al. Analysis of equid herpesvirus 1 strain variation reveals a point mutation of the DNA polymerase strongly associated with neuropathogenic versus nonneuropathogenic disease outbreaks. J Virol. 2006;80(8):4047–60.PubMedPubMedCentralCrossRef Nugent J, Birch-Machin I, Smith KC, Mumford JA, Swann Z, Newton JR, et al. Analysis of equid herpesvirus 1 strain variation reveals a point mutation of the DNA polymerase strongly associated with neuropathogenic versus nonneuropathogenic disease outbreaks. J Virol. 2006;80(8):4047–60.PubMedPubMedCentralCrossRef
21.
Zurück zum Zitat Kolb AW, Lewin AC, Moeller Trane R, McLellan GJ, Brandt CR. Phylogenetic and recombination analysis of the herpesvirus genus varicellovirus. BMC Genom. 2017;18(1):887.CrossRef Kolb AW, Lewin AC, Moeller Trane R, McLellan GJ, Brandt CR. Phylogenetic and recombination analysis of the herpesvirus genus varicellovirus. BMC Genom. 2017;18(1):887.CrossRef
22.
Zurück zum Zitat Vaz PK, Horsington J, Hartley CA, Browning GF, Ficorilli NP, Studdert MJ, et al. Evidence of widespread natural recombination among field isolates of equine herpesvirus 4 but not among field isolates of equine herpesvirus 1. J Gen Virol. 2016;97(3):747–55.PubMedPubMedCentralCrossRef Vaz PK, Horsington J, Hartley CA, Browning GF, Ficorilli NP, Studdert MJ, et al. Evidence of widespread natural recombination among field isolates of equine herpesvirus 4 but not among field isolates of equine herpesvirus 1. J Gen Virol. 2016;97(3):747–55.PubMedPubMedCentralCrossRef
23.
Zurück zum Zitat Diallo IS, Hewitson G, Wright L, Rodwell BJ, Corney BG. Detection of equine herpesvirus type 1 using a real-time polymerase chain reaction. J Virol Methods. 2006;131(1):92–8.PubMedCrossRef Diallo IS, Hewitson G, Wright L, Rodwell BJ, Corney BG. Detection of equine herpesvirus type 1 using a real-time polymerase chain reaction. J Virol Methods. 2006;131(1):92–8.PubMedCrossRef
24.
Zurück zum Zitat Lewin AC, Coghill LM, Mironovich M, Liu CC, Carter RT, Ledbetter EC. Phylogenomic analysis of global isolates of canid alphaherpesvirus 1. Viruses. 2020;12(12):1421.PubMedPubMedCentralCrossRef Lewin AC, Coghill LM, Mironovich M, Liu CC, Carter RT, Ledbetter EC. Phylogenomic analysis of global isolates of canid alphaherpesvirus 1. Viruses. 2020;12(12):1421.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Marino ME, Mironovich MA, Ineck NE, Citino SB, Emerson JA, Maggs DJ, et al. Full viral genome sequencing and phylogenomic analysis of feline herpesvirus type 1 (FHV-1) in cheetahs (Acinonyx jubatus). Viruses. 2021;13(11):2307.PubMedPubMedCentralCrossRef Marino ME, Mironovich MA, Ineck NE, Citino SB, Emerson JA, Maggs DJ, et al. Full viral genome sequencing and phylogenomic analysis of feline herpesvirus type 1 (FHV-1) in cheetahs (Acinonyx jubatus). Viruses. 2021;13(11):2307.PubMedPubMedCentralCrossRef
26.
Zurück zum Zitat Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.PubMedPubMedCentralCrossRef Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.PubMedPubMedCentralCrossRef
27.
Zurück zum Zitat Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14(6):587–9.PubMedPubMedCentralCrossRef Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14(6):587–9.PubMedPubMedCentralCrossRef
28.
Zurück zum Zitat Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol. 2020;37(5):1530–4.PubMedPubMedCentralCrossRef Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol. 2020;37(5):1530–4.PubMedPubMedCentralCrossRef
29.
Zurück zum Zitat Huson DH, Bryant D. Application of phylogenetic networks in evolutionary studies. Mol Biol Evol. 2006;23(2):254–67.PubMedCrossRef Huson DH, Bryant D. Application of phylogenetic networks in evolutionary studies. Mol Biol Evol. 2006;23(2):254–67.PubMedCrossRef
30.
Zurück zum Zitat Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9.PubMedPubMedCentralCrossRef Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9.PubMedPubMedCentralCrossRef
31.
Zurück zum Zitat Martin DP, Murrell B, Golden M, Khoosal A, Muhire B. RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evol. 2015;1(1):vev003.PubMedPubMedCentralCrossRef Martin DP, Murrell B, Golden M, Khoosal A, Muhire B. RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evol. 2015;1(1):vev003.PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat Jin L, Nei M. Limitations of the evolutionary parsimony method of phylogenetic analysis. Mol Biol Evol. 1990;7(1):82–102.PubMed Jin L, Nei M. Limitations of the evolutionary parsimony method of phylogenetic analysis. Mol Biol Evol. 1990;7(1):82–102.PubMed
33.
Zurück zum Zitat Martin D, Rybicki E. RDP: detection of recombination amongst aligned sequences. Bioinformatics. 2000;16(6):562–3.PubMedCrossRef Martin D, Rybicki E. RDP: detection of recombination amongst aligned sequences. Bioinformatics. 2000;16(6):562–3.PubMedCrossRef
34.
Zurück zum Zitat Padidam M, Sawyer S, Fauquet CM. Possible emergence of new geminiviruses by frequent recombination. Virology. 1999;265(2):218–25.PubMedCrossRef Padidam M, Sawyer S, Fauquet CM. Possible emergence of new geminiviruses by frequent recombination. Virology. 1999;265(2):218–25.PubMedCrossRef
35.
36.
Zurück zum Zitat Posada D, Crandall KA. Evaluation of methods for detecting recombination from DNA sequences: computer simulations. Proc Natl Acad Sci U S A. 2001;98(24):13757–62.PubMedPubMedCentralCrossRef Posada D, Crandall KA. Evaluation of methods for detecting recombination from DNA sequences: computer simulations. Proc Natl Acad Sci U S A. 2001;98(24):13757–62.PubMedPubMedCentralCrossRef
37.
Zurück zum Zitat Gibbs MJ, Armstrong JS, Gibbs AJ. Sister-scanning: a Monte Carlo procedure for assessing signals in recombinant sequences. Bioinformatics. 2000;16(7):573–82.PubMedCrossRef Gibbs MJ, Armstrong JS, Gibbs AJ. Sister-scanning: a Monte Carlo procedure for assessing signals in recombinant sequences. Bioinformatics. 2000;16(7):573–82.PubMedCrossRef
38.
Zurück zum Zitat Lewin AC, Kolb AW, McLellan GJ, Bentley E, Bernard KA, Newbury SP, et al. Genomic, recombinational and phylogenetic characterization of global feline herpesvirus 1 isolates. Virology. 2018;518:385–97.PubMedCrossRef Lewin AC, Kolb AW, McLellan GJ, Bentley E, Bernard KA, Newbury SP, et al. Genomic, recombinational and phylogenetic characterization of global feline herpesvirus 1 isolates. Virology. 2018;518:385–97.PubMedCrossRef
39.
Zurück zum Zitat Malik P, Bálint A, Dán A, Pálfi V. Molecular characterisation of the ORF68 region of equine herpesvirus-1 strains isolated from aborted fetuses in Hungary between 1977 and 2008. Acta Vet Hung. 2012;60(1):175–87.PubMedCrossRef Malik P, Bálint A, Dán A, Pálfi V. Molecular characterisation of the ORF68 region of equine herpesvirus-1 strains isolated from aborted fetuses in Hungary between 1977 and 2008. Acta Vet Hung. 2012;60(1):175–87.PubMedCrossRef
40.
Zurück zum Zitat Matczuk AK, Skarbek M, Jackulak NA, Bażanów BA. Molecular characterisation of equid alphaherpesvirus 1 strains isolated from aborted fetuses in Poland. Virol J. 2018;15(1):186.PubMedPubMedCentralCrossRef Matczuk AK, Skarbek M, Jackulak NA, Bażanów BA. Molecular characterisation of equid alphaherpesvirus 1 strains isolated from aborted fetuses in Poland. Virol J. 2018;15(1):186.PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Lunn DP, Davis-Poynter N, Flaminio MJ, Horohov DW, Osterrieder K, Pusterla N, et al. Equine herpesvirus-1 consensus statement. J Vet Intern Med. 2009;23(3):450–61.PubMedCrossRef Lunn DP, Davis-Poynter N, Flaminio MJ, Horohov DW, Osterrieder K, Pusterla N, et al. Equine herpesvirus-1 consensus statement. J Vet Intern Med. 2009;23(3):450–61.PubMedCrossRef
42.
Zurück zum Zitat Mori E, Borges AS, Delfiol DJ, Oliveira Filho JP, Gonçalves RC, Cagnini DQ, et al. First detection of the equine herpesvirus 1 neuropathogenic variant in Brazil. Rev Sci Tech. 2011;30(3):949–54.PubMedCrossRef Mori E, Borges AS, Delfiol DJ, Oliveira Filho JP, Gonçalves RC, Cagnini DQ, et al. First detection of the equine herpesvirus 1 neuropathogenic variant in Brazil. Rev Sci Tech. 2011;30(3):949–54.PubMedCrossRef
43.
Zurück zum Zitat Smith KL, Allen GP, Branscum AJ, Frank Cook R, Vickers ML, Timoney PJ, et al. The increased prevalence of neuropathogenic strains of EHV-1 in equine abortions. Vet Microbiol. 2010;141(1–2):5–11.PubMedCrossRef Smith KL, Allen GP, Branscum AJ, Frank Cook R, Vickers ML, Timoney PJ, et al. The increased prevalence of neuropathogenic strains of EHV-1 in equine abortions. Vet Microbiol. 2010;141(1–2):5–11.PubMedCrossRef
44.
Zurück zum Zitat Pusterla N, Barnum S, Lawton K, Wademan C, Corbin R, Hodzic E. Investigation of the EHV-1 genotype (N(752), D(752), and H(752)) in swabs collected from equids with respiratory and neurological disease and abortion from the United States (2019–2022). J Equine Vet Sci. 2023;123: 104244.PubMedCrossRef Pusterla N, Barnum S, Lawton K, Wademan C, Corbin R, Hodzic E. Investigation of the EHV-1 genotype (N(752), D(752), and H(752)) in swabs collected from equids with respiratory and neurological disease and abortion from the United States (2019–2022). J Equine Vet Sci. 2023;123: 104244.PubMedCrossRef
45.
Zurück zum Zitat Pusterla N, Barnum S, Miller J, Varnell S, Dallap-Schaer B, Aceto H, et al. Investigation of an EHV-1 outbreak in the United States Caused by a new H(752) genotype. Pathogens. 2021;10(6):747.PubMedPubMedCentralCrossRef Pusterla N, Barnum S, Miller J, Varnell S, Dallap-Schaer B, Aceto H, et al. Investigation of an EHV-1 outbreak in the United States Caused by a new H(752) genotype. Pathogens. 2021;10(6):747.PubMedPubMedCentralCrossRef
46.
Zurück zum Zitat Sutton G, Thieulent C, Fortier C, Hue ES, Marcillaud-Pitel C, Pléau A, et al. Identification of a new equid herpesvirus 1 DNA polymerase (ORF30) genotype with the isolation of a C(2254)/H(752) strain in French horses showing no major impact on the strain behaviour. Viruses. 2020;12(10):1160.PubMedPubMedCentralCrossRef Sutton G, Thieulent C, Fortier C, Hue ES, Marcillaud-Pitel C, Pléau A, et al. Identification of a new equid herpesvirus 1 DNA polymerase (ORF30) genotype with the isolation of a C(2254)/H(752) strain in French horses showing no major impact on the strain behaviour. Viruses. 2020;12(10):1160.PubMedPubMedCentralCrossRef
47.
Zurück zum Zitat Loncoman CA, Vaz PK, Coppo MJ, Hartley CA, Morera FJ, Browning GF, et al. Natural recombination in alphaherpesviruses: Insights into viral evolution through full genome sequencing and sequence analysis. Infect Genet Evol. 2017;49:174–85.PubMedCrossRef Loncoman CA, Vaz PK, Coppo MJ, Hartley CA, Morera FJ, Browning GF, et al. Natural recombination in alphaherpesviruses: Insights into viral evolution through full genome sequencing and sequence analysis. Infect Genet Evol. 2017;49:174–85.PubMedCrossRef
48.
Zurück zum Zitat Burrel S, Boutolleau D, Ryu D, Agut H, Merkel K, Leendertz FH, et al. Ancient recombination events between human herpes simplex viruses. Mol Biol Evol. 2017;34(7):1713–21.PubMedPubMedCentralCrossRef Burrel S, Boutolleau D, Ryu D, Agut H, Merkel K, Leendertz FH, et al. Ancient recombination events between human herpes simplex viruses. Mol Biol Evol. 2017;34(7):1713–21.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Kolb AW, Larsen IV, Cuellar JA, Brandt CR. Genomic, phylogenetic, and recombinational characterization of herpes simplex virus 2 strains. J Virol. 2015;89(12):6427–34.PubMedPubMedCentralCrossRef Kolb AW, Larsen IV, Cuellar JA, Brandt CR. Genomic, phylogenetic, and recombinational characterization of herpes simplex virus 2 strains. J Virol. 2015;89(12):6427–34.PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Lee K, Kolb AW, Sverchkov Y, Cuellar JA, Craven M, Brandt CR. Recombination analysis of herpes simplex virus 1 reveals a bias toward GC content and the inverted repeat regions. J Virol. 2015;89(14):7214–23.PubMedPubMedCentralCrossRef Lee K, Kolb AW, Sverchkov Y, Cuellar JA, Craven M, Brandt CR. Recombination analysis of herpes simplex virus 1 reveals a bias toward GC content and the inverted repeat regions. J Virol. 2015;89(14):7214–23.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Kuny CV, Szpara ML. Alphaherpesvirus genomics: past, present and future. Curr Issues Mol Biol. 2021;42:41–80.PubMed Kuny CV, Szpara ML. Alphaherpesvirus genomics: past, present and future. Curr Issues Mol Biol. 2021;42:41–80.PubMed
Metadaten
Titel
Phylogenomic assessment of 23 equid alphaherpesvirus 1 isolates obtained from USA-based equids
verfasst von
Ugochi Emelogu
Andrew C. Lewin
Udeni B. R. Balasuriya
Chin-Chi Liu
Rebecca P. Wilkes
Jianqiang Zhang
Erinn P. Mills
Renee T. Carter
Publikationsdatum
01.12.2023
Verlag
BioMed Central
Erschienen in
Virology Journal / Ausgabe 1/2023
Elektronische ISSN: 1743-422X
DOI
https://doi.org/10.1186/s12985-023-02248-z

Weitere Artikel der Ausgabe 1/2023

Virology Journal 1/2023 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Battle of Experts: Sport vs. Spritze bei Adipositas und Typ-2-Diabetes

11.05.2024 DDG-Jahrestagung 2024 Kongressbericht

Im Battle of Experts traten zwei Experten auf dem Diabeteskongress gegeneinander an: Die eine vertrat die Auffassung „Sport statt Spritze“ bei Adipositas und Typ-2-Diabetes, der andere forderte „Spritze statt Sport!“ Am Ende waren sie sich aber einig: Die Kombination aus beidem erzielt die besten Ergebnisse.

Vorsicht, erhöhte Blutungsgefahr nach PCI!

10.05.2024 Koronare Herzerkrankung Nachrichten

Nach PCI besteht ein erhöhtes Blutungsrisiko, wenn die Behandelten eine verminderte linksventrikuläre Ejektionsfraktion aufweisen. Das Risiko ist umso höher, je stärker die Pumpfunktion eingeschränkt ist.

Triglyzeridsenker schützt nicht nur Hochrisikopatienten

10.05.2024 Hypercholesterinämie Nachrichten

Patienten mit Arteriosklerose-bedingten kardiovaskulären Erkrankungen, die trotz Statineinnahme zu hohe Triglyzeridspiegel haben, profitieren von einer Behandlung mit Icosapent-Ethyl, und zwar unabhängig vom individuellen Risikoprofil.

Gibt es eine Wende bei den bioresorbierbaren Gefäßstützen?

In den USA ist erstmals eine bioresorbierbare Gefäßstütze – auch Scaffold genannt – zur Rekanalisation infrapoplitealer Arterien bei schwerer PAVK zugelassen worden. Das markiert einen Wendepunkt in der Geschichte dieser speziellen Gefäßstützen.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.