Skip to main content
Journal of Virology logoLink to Journal of Virology
. 2012 Apr;86(7):3617–3625. doi: 10.1128/JVI.07108-11

The Genome Sequence of the Emerging Common Midwife Toad Virus Identifies an Evolutionary Intermediate within Ranaviruses

Carla Mavian a, Alberto López-Bueno a, Ana Balseiro b, Rosa Casais b, Antonio Alcamí a, Alí Alejo c,*,
PMCID: PMC3302492  PMID: 22301140

Abstract

Worldwide amphibian population declines have been ascribed to global warming, increasing pollution levels, and other factors directly related to human activities. These factors may additionally be favoring the emergence of novel pathogens. In this report, we have determined the complete genome sequence of the emerging common midwife toad ranavirus (CMTV), which has caused fatal disease in several amphibian species across Europe. Phylogenetic and gene content analyses of the first complete genomic sequence from a ranavirus isolated in Europe show that CMTV is an amphibian-like ranavirus (ALRV). However, the CMTV genome structure is novel and represents an intermediate evolutionary stage between the two previously described ALRV groups. We find that CMTV clusters with several other ranaviruses isolated from different hosts and locations which might also be included in this novel ranavirus group. This work sheds light on the phylogenetic relationships within this complex group of emerging, disease-causing viruses.

INTRODUCTION

Global population declines and extinction of multiple amphibian species have been reported over the last 20 years (11). As estimated by the International Union for Conservation of Nature, the Amphibia class includes the highest number of critically endangered species among the animal classes examined, with an estimated 41% of its species under threat (23). Therefore, this class seems to be a particularly sensitive indicator of the current biodiversity losses associated with the global warming process and other human-related environmental factors. It has been shown that habitat loss, increased human population density, and increased climatic variability are important factors that compromise the survival of amphibian species (34) and that multiple drivers of extinction are likely to coordinately accelerate amphibian declines in the near future (20).

The role of emerging infectious diseases in the rapid decline of amphibian populations is being increasingly studied. The recent spread of the lethal fungal pathogen Batrachochytrium dendrobatidis is well documented and has been associated with long-term amphibian population declines in several locations (6). A second widespread pathogen of amphibians that is now recognized as an emerging infectious disease and therefore likewise included in the list of notifiable diseases by the World Organization for Animal Health is ranaviruses. Although their association with population declines has to be studied further (14), ranavirus infections have been clearly associated with mass mortalities of several amphibian as well as reptile and fish species worldwide (3, 36, 40).

Ranavirus is a genus within the Iridoviridae family which includes large, icosahedral viruses containing circular, double-stranded DNA genomes with sizes ranging from 105 kbp to 140 kbp and a coding potential of approximately 100 open reading frames (ORFs) (10). The family includes five different genera, two of which infect insects while the others infect cold-blooded vertebrates. Species from the genera Megalocytivirus and Lymphocystivirus have been found to infect only teleost fish, while ranaviruses are known to infect reptiles, amphibians, and fish, and the reasons for this broad host specificity are yet unknown.

Ranavirus outbreaks have been described from different locations worldwide. However, full-length genome sequences have been published only for Asian, American, and Australian isolates (18, 21, 24, 28, 35, 37, 41). Recently, it has been proposed that ranaviruses can be subdivided into two distinct groups based on phylogenetic analyses and genome colinearity, grouper iridovirus (GIV)-like ranaviruses and amphibian-like ranaviruses (ALRV) (24). The first group includes GIV and Singapore grouper iridovirus (SGIV), which were isolated in Asia from fish (35, 41). The ALRVs include frog virus 3 (FV3) and Ambystoma tigrinum virus (ATV), which were isolated from frogs and salamanders, respectively, in North America, the tiger frog virus (TFV), isolated in China, the epizootic hematopoietic necrosis virus (EHNV), isolated from fish in Australia, and the soft-shelled turtle iridovirus (STIV), isolated in China (21). Within the ALRVs, the degree of genome sequence colinearity is high, although two different groups of viruses can be distinguished: the EHNV/ATV group, which may be closer to a putative most recent common ancestor of the whole group (24), and the FV3/TFV/STIV group, which shows two discrete acquired genomic inversions when compared to the genomes of the former group.

Ranavirus emergence as a pathogen of amphibians and other ectothermic vertebrates is probably linked to their host range plasticity as well as to environmental and ecological factors (17). The first clear evidence that viral infection can be the cause of localized amphibian population declines was reported in Europe, where ranaviruses have caused large-scale mortalities of the common frog Rana temporaria in the United Kingdom since 1985 (40). The causative agent is thought to be a variant of FV3, which may have been introduced through the movement of infected animals from North America (22).

The common midwife toad virus (CMTV) was first isolated on the European continental mainland in 2007 from diseased tadpoles of the common midwife toad (Alytes obstetricans) in a high-altitude permanent water trough in the Picos de Europa National Park in Spain (4). The virus, causing a high mortality rate in this species as well as in juvenile alpine newts in the 2008 outbreak (Mesotriton alpestris cyreni) (5), was shown to be responsible for a systemic hemorrhagic disease. Common histological findings were the presence of intracytoplasmic inclusion bodies and the necrosis of endothelial cells, the latter of which results in destruction of several organs, including skin, liver, and kidney. Sequence analyses of DNA fragments belonging to the major capsid protein (MCP) and DNA polymerase genes showed that CMTV clustered more closely with the ALRVs than with the GIV-like ranaviruses within the Ranavirus genus. In 2010, a CMTV outbreak in a pond in the Netherlands was described as the cause of a mass mortality event affecting water frogs and common newts (29), showing that both the host range and geographic distribution of CMTV are much wider than previously suspected.

The importance of ranavirus infections in amphibian population declines as well as the lack of knowledge about the nature of circulating European ranaviruses prompted us to determine the complete genome sequence of the emerging CMTV.

MATERIALS AND METHODS

Virus and cells.

CMTV isolated from alpine newts (Mesotriton alpestris cyreni) without plaque purification (4) was amplified in a single step on zebrafish ZF4 cells (ATCC CRL-2050) at 28°C in Dutch-modified RPMI medium containing 2% fetal calf serum. Supernatants from CMTV-infected ZF4 cells were collected, and viral particles were obtained by ultracentrifugation at 20,000 × g through a 36% sucrose cushion for 1 h at 4°C.

Isolation of viral DNA.

The virus particles were treated with DNase I and S7 micrococcal nuclease to digest free DNA. After proteinase K treatment, viral DNA was extracted with phenol-chloroform and precipitated with sodium acetate and ethanol in the presence of 10 μg of glycogen (Roche) as the carrier. Extracted DNA was randomly amplified using Phi29 DNA polymerase (GenomiPhi V2; GE Healthcare).

Sequencing and assembly of the viral genome.

Amplified viral DNA was pyrosequenced on a 454 GS FLX instrument housed in the Parque Científico de Madrid. The output consists of 40,271 sequences with an average size of 375 bp. Reads were assembled in two steps with Newbler 2.5.3 (Roche-454 Life Science). First, a de novo assembly under stringent parameters (97% minimum overlap identity in at least 250 bp) generated 2 contigs flanked by repeated sequences. Then a unique contig was obtained by aligning the reads to the de novo genome under more relaxed parameters (98% identity in at least 50 bp). PCR amplification and Sanger sequencing were carried out in order to define homopolymer sequencing ambiguities at positions 73,736 and 3,392 and a polymorphism at position 11,812 as well as the precise number of TGTGAAGCGTAAGTCCCC repeats at position 66,494. The number of repetitions of the microsatellite region detected at position 38,092 was determined by running a specific PCR product in a 4% agarose gel (see Fig. 2). Sequences of primers are available upon request.

Fig 2.

Fig 2

PCR amplification of the microsatellites in CMTV and FV3. The microsatellites of FV3 (lane 2) and CMTV (lane 3) were PCR amplified using specific primers and run on a 4% agarose gel together with reference PCR products of known sizes (lane 4, 170 bp; lane 5, 174 bp; lane 6, 184 bp). Lanes 1 and 7, molecular weight markers.

Genome analysis and annotation.

Open reading frames (ORFs) were numbered consecutively from the same arbitrary start point as in ATV and EHNV (24, 28), and transcriptional sense was indicated by R (right) or L (left). The genome was annotated with the genome annotation transfer utility (GATU) software (39) using the genome of STIV as a template and further refined manually by using the similarity search algorithm BLASTP (http://blast.ncbi.nlm.nih.gov/) on all unassigned ORFs longer than 120 bp. Overlapping ORFs were annotated only if they had been previously annotated in other ranaviral genomes, preserving the same number for both R and L orientations (CMTV ORFs 35 and 54). Nucleotide-to-nucleotide comparisons between the CMTV genome and other iridoviral genomes as well as among different ranaviruses were calculated using the JDotter (Java Dot Plot Alignments) software with the default settings. The algorithm employed by this software is described in detail in reference 9. In the dot plots generated, a straight line represents a stretch of similarity between the two sequences compared on the x and y axes. The full-length genome sequences (accession numbers) used in the dot plot analyses were FV3 (AY548484), TFV (AF389451), ATV (AY150217), EHNV (FJ433873), GIV (AY666015), SGIV (AY521625), and STIV (NC012637). Phylogenetic trees were constructed using MEGA5 software. For the phylogenetic analysis of ranavirus MCPs, the following sequences were used: ranavirus KRV-1 (KRV-1; accession no. ADO14139), Rana catesbeiana virus-JP (RCV-JP; no. BAH80413), soft-shelled turtle iridovirus (STIV; no. ABC59813), frog virus 3 (FV3; no. ACP19256), tiger frog virus (TFV; no. AAK55105), Bohle iridovirus (BIV; no. ACO90022), pike-perch iridovirus (PPIV; no. ACO90019), Rana esculenta virus (REV; no. ACO90020), Chinese giant salamander virus (CGSV; no. ADZ47908), Ambystoma tigrinum virus (ATV; no. TP003785), epizootic hematopoietic necrosis virus (EHNV; no. AAO32315), Ranavirus maxima/9995205/DNK (Rmax; no. ADI71344), cod iridovirus/15/04.11.92/DNK (CodV; no. ADI71343), European sheatfish virus (ESV; no. ACO90018), European catfish virus (ECV; no. ACO90017), short-finned eel ranavirus (SERV; no. ACO90021), largemouth bass ulcerative syndrome virus (LMBUSV; no. ADB77863), largemouth bass virus (LMBV; no. CBW46836), grouper iridovirus (GIV; no. AEI85923), Singapore grouper iridovirus (SGIV; no. AAS18087), king grouper iridovirus (KGIV; no. AEI85924), crimson snapper iridovirus (CSIV; no. AEI85915), and lymphocystis disease virus from China (LCDV-C; no. YP_025102).

Nucleotide sequence accession number.

The genome sequence of CMTV was deposited into GenBank under accession no. JQ231222.

RESULTS

CMTV is a distinct ranavirus isolated in Europe.

To better understand the taxonomic position of the CMTV, we sequenced its MCP gene and performed a phylogenetic analysis comparing the MCP gene to that of 23 other ranaviruses. As shown in Fig. 1A, the CMTV clustered very closely with the Chinese giant salamander virus, isolated in China in 2010 (16), and the Rana esculenta virus, isolated from edible frogs (Pelophylax esculentus) in Italy and Denmark in 2009 (3). Together with the pike-perch iridovirus, isolated from pike-perch fry in Finland in 1998 (38), all four viruses formed a distinct group that was included within the previously described ALRVs but clearly separated from its two proposed major branches, represented by FV3 and EHNV. Collectively, the CMTV-like viruses are found to be more closely related to the FV3 group, suggesting both may have diverged from a common ancestor. Finally, the CMTV was distantly related to other ranaviruses, such as Ranavirus maxima/9995205/DNK, cod iridovirus/15/04.11.92/DNK (2), European sheatfish virus (1), and European catfish virus (32), that were isolated from fish on the European continent. As no full-length genome sequences for ranaviruses isolated in Europe have been described to date (Fig. 1B), and given the apparently distinct position of CMTV within the ALRVs, we decided to sequence its complete genome.

Fig 1.

Fig 1

CMTV is a novel amphibian-like ranavirus isolated in Europe. (A) Phylogenetic tree of ranavirus major capsid protein sequences. The sequences were aligned with ClustalW and trimmed manually to the 450-amino-acid sequence available for the Chinese giant salamander virus (CGSV) MCP, and phylogenetic reconstruction was obtained by neighbor-joining with 1,000 bootstrap replicates. The consensus bootstrap tree is shown, with confidence values indicated on the branches. Branches with bootstrap values below 50% are collapsed. Lymphocystis disease virus from China (LCDV-C) was used as the outgroup. Viruses infecting amphibian or reptile hosts are marked with an asterisk. The GenBank accession numbers of the Iridovirus MCP sequences used for this analysis are listed in Materials and Methods. (B) World map showing the locations of isolation and the hosts of the published full-length ranavirus genome sequences and CMTV.

CMTV is a typical ALRV.

The complete genomic sequence of the CMTV was obtained using pyrosequencing combined with Sanger sequencing of PCR products for unresolved regions. After assembly, we obtained a final genome size of 106,878 bp, with an average coverage of 128 reads per position. The genome size of CMTV is smaller than that of the EHNV (127,011 bp) and slightly larger than those of the other fully sequenced ALRVs ATV (106,332 bp), FV3 (105,903 bp), STIV (105,890 bp), and TFV (105,057 bp). The 55.3% GC content of the CMTV genome is similar to those of the other ALRVs, which range from 54% to 55%, but higher than those of the GIV (49%) and SGIV (48%) ranaviruses.

The microsatellite identified in FV3 (12) and STIV was also found in CMTV. The use of primers specific for the flanking sequences produced a larger PCR amplification product on CMTV DNA than on FV3 DNA, showing that the region in CMTV contains 60 dinucleotide repeats, compared to the 34 copies observed in the other two viruses, and that it can therefore be used for differentiation of these viruses (Fig. 2).

The CMTV genome was found to contain 104 ORFs (Fig. 3) encoding putatively expressed proteins with predicted molecular masses ranging from 5.2 kDa to 144.2 kDa and with conserved domains assigned to structural proteins as well as proteins potentially involved in replication, transcription, and host response modification (Table 1). For 101 CMTV-predicted proteins, orthologues were readily found in other ranavirus (24) genomes, with identities ranging from 65% to 100%. Only three ORFs (11L, 96L, and 100L) had not been previously annotated in other ranaviruses, but no significant similarities to other known proteins were found in the databases. Early stop codons or frameshift mutations account for the absence of annotated protein orthologues for these three ORFs in other ALRVs in spite of their relatively high conservation at the nucleotide level. Additionally, a genomic inversion in the FV3 group split the CMTV ORF 100L in these three ranaviruses. Whether these proteins are truly expressed and functional during CMTV infection remains to be addressed.

Fig 3.

Fig 3

Linear schematic organization of CMTV genome. Predicted open reading frames (ORFs) are represented as arrows indicating the size and direction of transcription. The black lines represent the CMTV genome and are divided into 30-kb segments. ORFs in the forward strand are drawn above the genome line and those in the complement strand are drawn below. Iridovirus core genes, ranavirus-specific genes, and amphibian-like ranavirus-specific genes are indicated in red, green, and blue, respectively. The asterisk shows the position of the microsatellite repetition.

Table 1.

Predicted CMTV open reading framesa

ORF Position (nt range) Product size (aa) Predicted function and conserved domain or signatureb Best BLASTP hitc
EHNV orthologued
FV3 orthologued
ORF % ID Accession no. ORF % ID ORF % ID
1R 16–786 256 Putative replication factor; pfam04947 FV3 1R 98 YP_031579 100R 98 1R 98
2L 1,497–2,510 337 Putative myristylated membrane protein STIV 2L 90 YP_002854233 1L 91 2L 90
3L 2,548–3,387 279 EHNV 2L 97 ACO25192 2L 97 NA
4R 3,419–4,633 404 TFV 4R 98 ABB92272.1 3R 97 3R 96
5R 4,674–4,856 60 FV3 4R 97 YP_031582.1 4R 92 4R 97
6R 5,291–5,893 200 cl03568 FV3 5R 86 YP_031583.1| NA NA 5R 86
7L 6,823–7,251 142 STIV 9L 95 YP_002854240.1 6L 78 NA
8R 7,246–11,211 1,321 Largest subunit of DNA-dependent RNA polymerase; cl11429 TFV 8R 99 AAL77794.1 7R 98 8R 98
9L 11,566–14,412 948 Helicase; smart00487 FV3 9L 98 YP_031587.1 8L 98 9L 98
10R 14,428–14,841 137 ATV p8 99 YP_003779.1 9R 97 10R 96
11L 15,222–15,455 77 NA NA
12L 15,499–16,593 364 Putative DNA repair protein; cl14812, cl14815 Rana grylio virus 9808 XPG 98 AAY43134.1 10L 97 95R 98
13R 16,689–17,156 155 pfam10881 FV3 94L 97 YP_031673.1 11R 97 94L 97
14R 17,217–17,432 71 STIV 98L 85 YP_002854329.1 12R 95 93L 96
15L 18,167–19,354 395 Immediate early protein ICP-46 STIV 97R 99 YP_002854328.1 13L 98 91R 98
16L 19,478–20,869 463 Major capsid protein; pfam04451 PPIV MCP 99 ACO90019.1 14L 99 90R 98
17L 20,962–22,041 359 STIV 95R 88 YP_002854326.1 15L 88 89R 86
18L 22,109–22,561 150 Thiol oxidoreductase; pfam04777 STIV 94R 100 YP_002854325.1 16L 97 88R 99
19R 22,594–24,393 599 TFV 93L 95 ABB92341.1 17R 94 87L 96
20R 24,746–24,961 71 TFV 92L 80 ABB92340.1 NA 86L 75
21L 25,406–25,993 195 Thymidine kinase; cd01673 TFV 91R 98 ABB92339.1 18L 96 85R 96
22L 26,068–26,805 245 Proliferating cell nuclear antigen EHNV 19L 99 ACO25209.1 19L 99 84R 98
23L 27,222–27,866 214 Cytosine DNA methyltransferase; cl12011 FV3 83R 99 YP_031662.1 20L 99 83R 99
24L 28,233–28,520 95 Thymidylate synthase; cl00358 EHNV 21L 94 ACO25211.1 21L 94 NA
25L 28,816–29,289 157 Putative immediate early protein; cl12687 EHNV 22L 97 ACO25212.1| 22L 97 82R 93
26L 29,418–29,696 92 Transcription elongation factor SII; cl02609 FV3 81R 97 YP_031660.1 23L 97 81R 97
27R 29,752–30,870 372 RNase III; cd00593, TIGR02191 EHNV 24R 99 ACO25214.1 24R 99 80L 99
28L 31,495–33,213 572 STIV 86R 95 YP_002854317.1 25L 83 79R 95
29R 33,298–33,987 229 STIV 85L 94 YP_002854316.1 26R 95 78L 95
30R 34,687–35,034 115 TFV 82L 99 ABB92335.1 27R 96 77L 98
31L 35,031–35,252 73 TFV 81R 99 ABB92334.1 28L 96 76R 96
32R 35,315–35,569 84 LITAF-like protein; cl02754 FV3 75L 100 YP_031654.1 29R 99 75L 100
33R 35,626–36,807 393 EHNV 30R 97 ACO25220.1 30R 97 74L 85
34R 36,947–37,999 350 Putative NTPase/helicase EHNV 31R 98 ACO25221.1 31R 98 73L 97
35R 38,496–39,212 238 TFV 77L 98 ABB92331.1 32R 100 72L 97
35L 38,502–38,984 160 EHNV 33L 96 ACO25223.1 33L 96 NA
36L 39,269–39,502 77 FV3 71R 96 YP_031650.1 34L 90 71R 96
37L 39,542–39,916 124 FV3 70R 98 YP_031649.1 35L 97 70R 98
38L 39,934–40,200 88 FV3 69R 100 YP_031648.1 36L 99 69R 100
39R 40,318–40,782 154 cl00060 EHNV 37R 97 ACO25227.1 37R 97 NA
40R 41,407–42,570 387 Ribonucleotide reductase, small subunit; cd01049 ATV p39 99 YP_003810.1 38R 98 67L 98
41R 42,625–42,978 117 STIV 70L 97 YP_002854301.1 39R 90 66L 91
42R 43,115–43,459 114 EHNV 40R 89 ACO25230.1 40R 89 65L 96
43L 43,852–44,139 95 CARD-like caspase; cl14633 STIV 67R 97 YP_002854298.1 41L 96 64R 93
44L 44,250–44,744 164 dUTPase; cd07557 STIV 66R 98 YP_002854297.1 42L 97 63R 96
45L 44,861–45,403 180 STIV 65R 92 YP_002854296.1 NA NA
46R 45,123–48,776 1,217 DNA-dependent RNA polymerase B subunit; cd00653, cl04593, COG0085 FV3 62L 98 YP_031641.1 43R 96 62L 98
47L 49,315–52,356 1,013 DNA polymerase; cl10023, cl10012 STIV 63R 99 YP_002854294.1 44L 99 60R 99
48R 52,518–53,576 352 FV3 59L 98 YP_031638.1 45R 97 59L 98
49L 54,062–54,616 184 EHNV 46L 98 ACO25236.1 46L 98 NA
50L 54,667–54,912 81 cl08321 EHNV 47L 88 ACO25237.1 47L 88 NA
51L 54,995–56,491 498 Putative phosphotransferase STIV 60R 99 YP_002854291.1 48L 98 57R 98
52L 56,532–56,936 134 EHNV 49L 98 ACO25239.1 49L 98 NA
53R 56,973–57,122 49 EHNV 50R 100 ACO25240.1 50R 100 NA
54R 57,130–58,425 431 Helicase; cd00046, COG1061 TFV 56L 98 AAL77803.1 51R 98 55L 97
54L 57,142–58,281 379 FV3 55R 97 YP_031634.1 p40 97 55R 97
55R 58,338–58,730 130 STIV 56L 98 YP_002854287.1 NA NA
56L 58,875–60,443 522 Myristylated membrane protein; pfam02442 FV3 53R 99 YP_031631.1 53L 98 53R 99
57R 60,780–61,847 355 3-beta-hydroxysteroid dehydrogenase; pfam01073 STIV 54L 99 YP_002854285.1 54R 97 52L 98
58L 62,104–63,789 561 TFV 53R 98 ABB92313.1 84R 98 51R 98
59R 63,868–65,379 503 cl02640, PTZ00108 STIV 52L 96 YP_002854283.1 83L 97 49L 98
60R 65,427–65,738 103 ATV p78 88 YP_003849.1 82L 98 48L 98
61R 65,741–66,157 138 TFV 49L 99 ABB92309.1 81L 96 47L 98
62R 66,282–66,749 155 Neurofilament triplet H1-like protein; PTZ00449 STIV 49L 94 YP_002854280.1 80L 83 46L 83
63R 66,859–67,269 136 FV3 45L 98 YP_031623.1 79L 98 45L 98
64R 67,396–68,364 322 TFV 46L 80 ABB92307.1 78L 96 42L 98
65L 68,753–72,331 1,192 FV3 41R 98 YP_031619.1 77R 98 41R 98
66L 72,321–72,461 46 ATV p71 91 YP_003842.1 76R 70 NA
67L 72,647–73,330 227 ATV p70 84 YP_003841.1 75R 80 40R 79
68L 73,346–73,696 116 EHNV 74R 97 ACO25264.1 74R 97 39R 94
69L 73,805–75,502 565 Ribonucleoside-diphosphate reductase, large subunit; cd01679, PRK09102 TFV 41R 99 AAL77800.1 73R 99 38R 98
70L 75,641–76,276 211 Putative NIF/NLI interacting factor; cl11391 EHNV 72R 99 ACO25262.1 72R 99 37R 98
71R 76,657–76,956 99 FV3 36L 89 YP_031614.1 NA 36L 89
72R 77,010–77,228 72 FV3 36L 93 YP_031614.1 NA 36L 93
73R 77,268–77,846 192 EHNV 71L 65 ACO25261.1 71L 65 35L 64
74L 77,831–78,151 106 EHNV 70R 94 ACO25260.1 70R 94 34R 95
75L 78,293–78,484 63 EHNV 69R 97 ACO25259.1 69R 97 33R 97
76L 78,567–80,717 716 Neurofilament triplet H1-like protein; cl06505, cl06430 STIV 35R 84 YP_002854266.1 68R 85 32R 81
77L 80,767–81,186 139 cl10444 EHNV 67R 98 ACO25257.1 67R 98 31R 95
78L 81,674–81,940 88 EHNV 66R 100 ACO25256.1 66R 100 NA
79L 82,077–82,565 162 STIV 32R 99 YP_002854263.1 63R 92 28R 98
80L 82,614–85,544 976 Putative tyrosine kinase; cl14933 STIV 31R 98 YP_002854262.1 62R 96 27R 97
81L 86,043–86,855 270 eIF2a-like protein; cl09927 EHNV 61R 95 ACO25251.1 61R 95 26R 83
82L 87,180–88,094 304 p31K protein; cl12506 EHNV 60R 97 ACO25250.1 60R 97 25R 99
83L 88,158–89,255 365 FV3 24R 98 YP_031602.1 57R 96 24R 98
84L 89,680–90,828 382 FV3 23R 98 YP_031601.1 56R 97 23R 98
85L 91,206–94,133 975 Putative D5 family NTP/ATPase; cl11759, cl07361, cl07360 STIV 25R 99 YP_002854256.1 85L 98 22R 99
86R 94,258–94,926 222 ATV p82 95 YP_003853.1 86R 95 21L 95
87L 95,163–95,609 148 STIV 23R 99 YP_002854254.1 88L 98 20R 96
88L 95,657–98,443 928 cl07414 TFV 19R 96 ABB92284.1 89L 88 19R 94
89R 98,232–98,468 78 FV3 18L 97 YP_031596.1 NA 18L 97
90R 98,505–100,013 502 STIV 18L 99 YP_002854249.1 90R 99 17L 99
91L 99,224–99,730 168 STIV 19R 99 YP_002854250.1 NA NA
92R 100,050–100,997 315 STIV 17L 97 YP_002854248.1 91R 98 NA
93L 100,254–101,081 275 Putative integrase homologue FV3 16R 99 YP_031594.1 NA 16R 99
94L 101,361–102,308 315 Putative A32-like virion packaging ATPase; cl04659, smart00382 STIV 16R 99 YP_002854247.1 92L 97 15R 97
95L 102,405–102,764 119 STIV 15R 100 YP_002854246.1 93L 98 14R 100
96L 102,841–103,035 64 NA NA
97R 103,076–103,270 64 TFV 13L ABB92279.1 NA NA
98R 103,382–104,425 347 STIV 14L 98 YP_002854245.1 95R 97 12L 98
99L 104,491–104,703 70 STIV 13R 99 YP_002854244.1 96L 99 11R 97
100L 104,769–104,990 73 NA NA
101R 105,187–105,873 228 EHNV 98R 97 ACO25288.1 98R 97 96R 94
102R 105,939–106,352 137 STIV 103R 99 YP_002854334.1 99R 96 97R 99
a

nt, nucleotides; aa, amino acids; ID, identity.

b

Predicted function is based on conserved domains and/or previous annotation in other ranaviruses. LITAF, lipopolysaccharide-induced tumor necrosis factor alpha (TNF-α).

c

Significant hits using conserved domain search at NCBI BLASTP are shown.

d

NA, not applicable (BLASTP E value > 0.001).

Analyses of all sequenced iridovirus genomes have identified 26 conserved genes, which make up the core iridovirus genes (15). As expected, these genes are conserved in CMTV. Additionally, 27 more genes were described to be conserved throughout the Ranavirus genus based on the EHNV gene content (24). These include a multigene family with five members in EHNV of which only two are retained in ATV and the FV3 group as well as in CMTV (ORFs 82L and 83L). Finally, the set of 13 genes conserved in all ALRVs (24) was also found in CMTV. Altogether, these results show that the CMTV can be classified as a typical ALRV in terms of gene content.

CMTV represents an intermediate in ALRV evolution.

Due to the >90% identity among ranavirus MCP sequences, phylogenetic studies based exclusively on these sequences may be insufficient for differentiating among ranavirus types. To further establish the phylogenetic relationships between CMTV and other fully sequenced ranaviruses, we performed an analysis of the concatenated protein sequences derived from the 26 iridoviral core genes (15). As previously reported, the Asian fish viruses GIV and SGIV formed a separated branch within the fully sequenced ranaviruses (Fig. 4). CMTV was found to cluster very confidently within the FV3-like group, which was clearly resolved from the EHNV-like group. Furthermore, CMTV sequences were found to have a slightly higher similarity to the FV3/STIV group than to TFV.

Fig 4.

Fig 4

CMTV is closely related to the FV3-like viruses. Phylogenetic analysis of the concatenated sequences of the 26 iridovirus core proteins (15) from the indicated viruses. The sequences were aligned with ClustalW, and the phylogenetic reconstruction was obtained by neighbor-joining with 1,000 bootstrap replicates. Consensus bootstrap confidence values are indicated above the branches. LCDV-C was used as the outgroup.

To determine the degree of colinearity of the CMTV genome with those of other ranaviruses, we performed dot plot analyses (Fig. 5 and data not shown). As expected, CMTV did not show major colinearity stretches when compared to the GIV-like fish ranaviruses (data not shown). As described before, ALRVs can be divided into two separate groups in which colinearity is maintained, one including EHNV and ATV and the second including FV3, STIV, and TFV. The CMTV genome was found to not be colinear with either of these groups. Specifically, compared to EHNV, an inversion of the CMTV genomic segment located between positions ∼62,000 and ∼91,000 was observed, while in comparison to FV3, a single different inversion affecting positions ∼15,000 and ∼105,000 was detected. These sites of genomic rearrangements correspond exactly to those of the double genomic inversion identified previously when FV3 was compared to EHNV (24). This suggests that CMTV or its ancestor may occupy an intermediate position in the evolutionary process that gave rise to the FV3/STIV/TFV group from the EHNV-like ancestor. Thus, inversion of one segment from an EHNV-like precursor may have produced a CMTV ancestor. A further inversion might have then produced the genomic structure found in the FV3/TFV group.

Fig 5.

Fig 5

CMTV is an evolutionary intermediate in ALRVs. Dot plot comparisons of the CMTV genome and the EHNV and FV3 genomes. For clarity, the reverse complement sequence of the FV3 genome was used in these comparisons. As a reference, a dot plot of EHNV versus FV3 is shown in the right panel. Black arrows indicate the sites of inversion in the CMTV genome compared to the FV3 genome, and gray arrows indicate the sites of inversion in the CMTV genome compared to the EHNV genome. Triangles indicate sequences present in CMTV but absent in FV3 or EHNV.

Upon closer inspection of the corresponding dot plots, CMTV was found to contain no major deletions and only three DNA sequence insertions of about 500 bp each compared to FV3 (Fig. 5). These insertions corresponded to sequences that are also found in EHNV and to different degrees in ATV and TFV but that are specifically lost in FV3 and STIV. In particular, the insertions affect CMTV 24L, CMTV 39R, and CMTV 81L, which encode a truncated thymidylate synthase protein, a putatively secreted protein containing an intact FGF domain, and an eIF2α-like protein, respectively. Conversely, compared to EHNV, CMTV presents at least 14 deletions and two insertions, which affect CMTV 6R as well as CMTV 71R and CMTV 72R, that are also found in TFV, FV3, and STIV. This pattern of sequence gain and loss among ALRVs is also consistent with the intermediate position of CMTV in the evolution of EHNV-like viruses toward FV3-like viruses, since CMTV retains EHNV-like characteristics which are lost in FV3-like viruses and has already acquired sequences found previously only in FV3-like viruses. Therefore, the genomic content and structure of CMTV may resemble those of the last common ancestor of TFV, STIV, and FV3 viruses.

DISCUSSION

In this report, we have shown that the CMTV, the first European ranavirus to be completely sequenced, should be classified as an ALRV. Based on our analyses, we propose that the CMTV might represent, in terms of gene content and genomic structure, a very recent ancestor of the previously described group of FV3-like viruses (24). As the CMTV forms a cluster with ranaviruses isolated from different species and locations worldwide, including fish and amphibian viruses, we believe that these may form a single, novel group of ALRVs. The information provided by this report should therefore be useful to specifically address the relationship of these viruses with CMTV and to further define the taxonomic positions of different ranavirus isolates, which have been poorly described at the genetic level.

One of the most outstanding features of ranaviruses is their wide host range, which includes anuran and urodele amphibians, reptiles, and teleost fish. It has been proposed that the most recent common ancestor of all ALRVs was most probably a fish virus, with independent events giving rise to isolates infecting different hosts (24). Experimental evidence suggests that ATV may infect only urodele hosts, while neither fish nor frogs are susceptible to ATV infection (25). As CMTV can infect both anuran and urodele amphibians in the wild (4, 5, 29), it is tempting to speculate that CMTV may have derived from an ATV-related virus, acquiring the ability to infect frogs. Potentially, further divergence into the FV3-like viruses may have produced frog- and reptile-specific viruses. However, FV3 infections of both fish and salamander species have been reported (30), and the Bohle iridovirus, isolated from the ornate burrowing frog in Australia, is also known to infect fish (31). Therefore, both data from experimental infections and a clearer taxonomic description of ranaviruses based on complete genome information will help address this issue.

As described above, CMTV has retained some features from the EHNV-like group which are lost in FV3-like viruses and has acquired potential coding sequences which were previously identified only in the FV3-like viruses, further supporting their intermediate position in the evolution between these groups of ALRVs. Interestingly, some of these coding regions may play a role in host range or virulence. In particular, a potentially functional protein belonging to the fibroblast growth factor family is present only in EHNV, ATV, and CMTV. In baculoviruses, such an activity was shown to be able to induce cell migration, suggesting a role in immunomodulation or viral spread (13). Additionally, the eIF2α-like protein (vIF2αH) present in most ALRVs, including CMTV, but truncated in the FV3 and STIV has been demonstrated to act as an inhibitor of the antiviral protein kinase PKR (33). Thus, ATV mutants lacking vIF2αH activity were shown to be more sensitive to the antiviral activity of interferon (IFN) and displayed an increased time to death in infected salamanders, demonstrating the protein's role as an important virulence factor in vivo (27). Conversely, CMTV 6R, which is found in ALRVs except EHNV and ATV, shows significant similarity to DNA and protein sequences from frog species, suggesting a recent acquisition event during the process of adaptation to novel amphibian hosts. Implications of these specific events of sequence gain and loss among ALRV groups on virus host range or pathogenicity remain to be addressed.

Whether the broad host range of ranaviruses, compared to that of other double-stranded DNA viruses like orthopoxviruses, is an inherent feature of the iridoviruses or related to a more recent evolutionary history of this virus family is an important topic for study. While gene loss has been established as one of the main driving forces in poxvirus evolution (19), we observe very little differences in terms of global gene content among ALRVs, which might indicate that this group of viruses is only on the verge of an evolutionary radiation. This may relate to the human-mediated exchange of host species or viruses between distant regions, which increases the number of potential hosts to infect and is consistent with their established character as emerging infectious-disease-causing agents.

As pathogens of wildlife, ranaviruses have been found to cause mass mortalities mainly in amphibian populations worldwide. One of the earliest amphibian population declines to be reported was that affecting the common midwife toad and other amphibian species, including salamanders, during the last decade in Spain (7, 8). These declines have been associated with the emergence of chytridiomycosis, a disease caused by the fungal pathogen Batrachochytrium dendrobatidis. Recent studies have shown the widespread yet heterogeneous distribution of Batrachochytrium dendrobatidis across the Iberian Peninsula and predict further spread of the disease in the future (42). However, the isolation of CMTV from diseased amphibians in Spain suggests that ranavirus infection, too, may be having an impact on population declines, as reported for other locations. In the Netherlands, animals affected with CMTV showed no evidence of Batrachochytrium dendrobatidis infection (29), suggesting that the two pathogens did not occur simultaneously. Whether an interaction between the pathogens is important for amphibian extinction dynamics remains to be addressed. Mass mortality events affecting the urodele Triturus marmoratus in northwest Portugal in 2003 are probably also related to ranavirus outbreaks, and recently, a novel ranavirus closely related to FV3 was isolated from lizards in Portugal (12). The nature of these viruses and whether they may represent different isolates of the CMTV are currently unknown. Although the distribution and degree of prevalence of CMTV in different host species need to be studied in greater detail, the presence of CMTV in the Netherlands (29) suggests that this virus may show a wide geographic distribution across the European mainland. In the case of tiger salamander infections in North America, it was shown that while ATV has coevolved with its host in their geographic range, human-mediated transfer of ATV-infected bait salamanders introduced novel, more virulent strains to discrete locations, resulting in disease emergence (26, 36). It will be important to determine the mechanisms of emergence of CMTV-caused disease and whether it may be linked to environmental factors, virus spread, or both.

As advances in sequencing technologies make more genomic information on the ranavirus isolates available, a better understanding of the evolutionary history of this virus family will be gained. In particular, a better knowledge of the ranavirus species and strain structure will be crucial to set up detection methods and surveillance strategies in order to minimize their impact on amphibian wildlife and cultivated species.

ACKNOWLEDGMENTS

This work was supported by grant AGL 2009-08711 from the Spanish Ministerio de Ciencia e Innovación. Alberto López-Bueno and Carla Mavián are recipients of the Ramón y Cajal and Formación del Personal Investigador fellowships, respectively, from the same institution.

Footnotes

Published ahead of print 1 February 2012

REFERENCES

  • 1. Ahne W, Schlotfeldt HJ, Thomsen I. 1989. Fish viruses: isolation of an icosahedral cytoplasmic deoxyribovirus from sheatfish (Silurus glanis). Zentralbl. Veterinarmed. B 36:333–336 [DOI] [PubMed] [Google Scholar]
  • 2. Ariel E, Holopainen R, Olesen NJ, Tapiovaara H. 2010. Comparative study of ranavirus isolates from cod (Gadus morhua) and turbot (Psetta maxima) with reference to other ranaviruses. Arch. Virol. 155:1261–1271 [DOI] [PubMed] [Google Scholar]
  • 3. Ariel E, et al. 2009. Ranavirus in wild edible frogs Pelophylax kl. esculentus in Denmark. Dis. Aquat. Organ. 85:7–14 [DOI] [PubMed] [Google Scholar]
  • 4. Balseiro A, et al. 2009. Pathology, isolation and molecular characterisation of a ranavirus from the common midwife toad Alytes obstetricans on the Iberian Peninsula. Dis. Aquat. Organ. 84:95–104 [DOI] [PubMed] [Google Scholar]
  • 5. Balseiro A, et al. 2010. Outbreak of common midwife toad virus in alpine newts (Mesotriton alpestris cyreni) and common midwife toads (Alytes obstetricans) in northern Spain: a comparative pathological study of an emerging ranavirus. Vet. J. 186:256–258 [DOI] [PubMed] [Google Scholar]
  • 6. Berger L, et al. 1998. Chytridiomycosis causes amphibian mortality associated with population declines in the rain forests of Australia and Central America. Proc. Natl. Acad. Sci. U. S. A. 95:9031–9036 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Bosch J, Martínez-Solano I. 2006. Chytrid fungus infection related to unusual mortalities of Salamandra salamandra and Bufo bufo in the Peñalara Natural Park, Spain. Oryx 40:84–89 [Google Scholar]
  • 8. Bosch J, Martínez-Solano I, García-París M. 2001. Evidence of a chytrid fungus infection involved in the decline of the common midwife toad (Alytes obstetricans) in protected areas of central Spain. Biol. Conserv. 97:331–337 [Google Scholar]
  • 9. Brodie R, Roper RL, Upton C. 2004. JDotter: a Java interface to multiple dotplots generated by dotter. Bioinformatics 20:279–281 [DOI] [PubMed] [Google Scholar]
  • 10. Chinchar VG, Hyatt A, Miyazaki T, Williams T. 2009. Family Iridoviridae: poor viral relations no longer. Curr. Top. Microbiol. Immunol. 328:123–170 [DOI] [PubMed] [Google Scholar]
  • 11. Collins JP. 2010. Amphibian decline and extinction: what we know and what we need to learn. Dis. Aquat. Organ. 92:93–99 [DOI] [PubMed] [Google Scholar]
  • 12. de Matos AP, et al. 2011. New viruses from Lacerta monticola (Serra da Estrela, Portugal): further evidence for a new group of nucleo-cytoplasmic large deoxyriboviruses. Microsc. Microanal. 17:101–108 [DOI] [PubMed] [Google Scholar]
  • 13. Detvisitsakun C, Berretta MF, Lehiy C, Passarelli AL. 2005. Stimulation of cell motility by a viral fibroblast growth factor homolog: proposal for a role in viral pathogenesis. Virology 336:308–317 [DOI] [PubMed] [Google Scholar]
  • 14. Duffus AL. 2009. Chytrid blinders: what other disease risks to amphibians are we missing? Ecohealth 6:335–339 [DOI] [PubMed] [Google Scholar]
  • 15. Eaton HE, et al. 2007. Comparative genomic analysis of the family Iridoviridae: re-annotating and defining the core set of iridovirus genes. Virol. J. 4:11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Geng Y, et al. 2011. First report of a ranavirus associated with morbidity and mortality in farmed Chinese giant salamanders (Andrias davidianus). J. Comp. Pathol. 145:95–102 [DOI] [PubMed] [Google Scholar]
  • 17. Gray MJ, Miller DL, Hoverman JT. 2009. Ecology and pathology of amphibian ranaviruses. Dis. Aquat. Organ. 87:243–266 [DOI] [PubMed] [Google Scholar]
  • 18. He JG, et al. 2002. Sequence analysis of the complete genome of an iridovirus isolated from the tiger frog. Virology 292:185–197 [DOI] [PubMed] [Google Scholar]
  • 19. Hendrickson RC, Wang C, Hatcher EL, Lefkowitz EJ. 2010. Orthopoxvirus genome evolution: the role of gene loss. Viruses 2:1933–1967 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Hof C, Araujo MB, Jetz W, Rahbek C. 2011. Additive threats from pathogens, climate and land-use change for global amphibian diversity. Nature 480:516–519 [DOI] [PubMed] [Google Scholar]
  • 21. Huang Y, et al. 2009. Complete sequence determination of a novel reptile iridovirus isolated from soft-shelled turtle and evolutionary analysis of Iridoviridae. BMC Genomics 10:224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Hyatt AD, et al. 2000. Comparative studies of piscine and amphibian iridoviruses. Arch. Virol. 145:301–331 [DOI] [PubMed] [Google Scholar]
  • 23. IUCN 2011. Numbers of threatened species by major groups of organisms (1996-2011), version 2011.2. http://www.iucnredlist.org/documents/summarystatistics/2011_2_RL_Stats_Table1.pdf
  • 24. Jancovich JK, Bremont M, Touchman JW, Jacobs BL. 2010. Evidence for multiple recent host species shifts among the ranaviruses (family Iridoviridae). J. Virol. 84:2636–2647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Jancovich JK, Davids EW, Seiler A, Jacobs BL, Collins JP. 2001. Transmission of the Ambystoma tigrinum virus to alternative hosts. Dis. Aquat. Organ. 46:159–163 [DOI] [PubMed] [Google Scholar]
  • 26. Jancovich JK, et al. 2005. Evidence for emergence of an amphibian iridoviral disease because of human-enhanced spread. Mol. Ecol. 14:213–224 [DOI] [PubMed] [Google Scholar]
  • 27. Jancovich JK, Jacobs BL. 2011. Innate immune evasion mediated by the Ambystoma tigrinum virus eukaryotic translation initiation factor 2α homologue. J. Virol. 85:5061–5069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Jancovich JK, et al. 2003. Genomic sequence of a ranavirus (family Iridoviridae) associated with salamander mortalities in North America. Virology 316:90–103 [DOI] [PubMed] [Google Scholar]
  • 29. Kik M, et al. 2011. Ranavirus-associated mass mortality in wild amphibians, the Netherlands, 2010: a first report. Vet. J. 190:284–286 [DOI] [PubMed] [Google Scholar]
  • 30. Mao J, Green DE, Fellers G, Chinchar VG. 1999. Molecular characterization of iridoviruses isolated from sympatric amphibians and fish. Virus Res. 63:45–52 [DOI] [PubMed] [Google Scholar]
  • 31. Moody NJG, Owens L. 1994. Experimental demonstration of the pathogenicity of a frog virus, Bohle iridovirus, for a fish species, barramundi-Lates-calcarifer. Dis. Aquat. Organ. 18:95–102 [Google Scholar]
  • 32. Pozet F, Morand M, Moussa A, Torhy C, de Kinkelin P. 1992. Isolation and preliminary characterization of a pathogenic icosahedral deoxyribovirus from the catfish Ictalurus melas. Dis. Aquat. Organ. 14:35–42 [Google Scholar]
  • 33. Rothenburg S, Chinchar VG, Dever TE. 2011. Characterization of a ranavirus inhibitor of the antiviral protein kinase PKR. BMC Microbiol. 11:56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Sodhi NS, et al. 2008. Measuring the meltdown: drivers of global amphibian extinction and decline. PLoS One 3:e1636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Song WJ, et al. 2004. Functional genomics analysis of Singapore grouper iridovirus: complete sequence determination and proteomic analysis. J. Virol. 78:12576–12590 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Storfer A, et al. 2007. Phylogenetic concordance analysis shows an emerging pathogen is novel and endemic. Ecol. Lett. 10:1075–1083 [DOI] [PubMed] [Google Scholar]
  • 37. Tan WG, Barkman TJ, Gregory Chinchar V, Essani K. 2004. Comparative genomic analyses of frog virus 3, type species of the genus Ranavirus (family Iridoviridae). Virology 323:70–84 [DOI] [PubMed] [Google Scholar]
  • 38. Tapiovaara H, Olesen NJ, Linden J, Rimaila-Parnanen E, von Bonsdorff CH. 1998. Isolation of an iridovirus from pike-perch Stizostedion lucioperca. Dis. Aquat. Organ. 32:185–193 [DOI] [PubMed] [Google Scholar]
  • 39. Tcherepanov V, Ehlers A, Upton C. 2006. Genome annotation transfer utility (GATU): rapid annotation of viral genomes using a closely related reference genome. BMC Genomics 7:150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Teacher AGF, Cunningham AA, Garner TWJ. 2010. Assessing the long-term impact of Ranavirus infection in wild common frog populations. Anim. Conserv. 13:514–522 [Google Scholar]
  • 41. Tsai CT, et al. 2005. Complete genome sequence of the grouper iridovirus and comparison of genomic organization with those of other iridoviruses. J. Virol. 79:2010–2023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Walker SF, et al. 2010. Factors driving pathogenicity vs. prevalence of amphibian panzootic chytridiomycosis in Iberia. Ecol. Lett. 13:372–382 [DOI] [PubMed] [Google Scholar]

Articles from Journal of Virology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES