Skip to main content
Journal of Virology logoLink to Journal of Virology
. 2009 Dec 30;84(6):2636–2647. doi: 10.1128/JVI.01991-09

Evidence for Multiple Recent Host Species Shifts among the Ranaviruses (Family Iridoviridae)

James K Jancovich 1, Michel Bremont 2, Jeffrey W Touchman 3, Bertram L Jacobs 1,3,*
PMCID: PMC2826071  PMID: 20042506

Abstract

Members of the genus Ranavirus (family Iridoviridae) have been recognized as major viral pathogens of cold-blooded vertebrates. Ranaviruses have been associated with amphibians, fish, and reptiles. At this time, the relationships between ranavirus species are still unclear. Previous studies suggested that ranaviruses from salamanders are more closely related to ranaviruses from fish than they are to ranaviruses from other amphibians, such as frogs. Therefore, to gain a better understanding of the relationships among ranavirus isolates, the genome of epizootic hematopoietic necrosis virus (EHNV), an Australian fish pathogen, was sequenced. Our findings suggest that the ancestral ranavirus was a fish virus and that several recent host shifts have taken place, with subsequent speciation of viruses in their new hosts. The data suggesting several recent host shifts among ranavirus species increase concern that these pathogens of cold-blooded vertebrates may have the capacity to cross numerous poikilothermic species barriers and the potential to cause devastating disease in their new hosts.


Iridoviruses are large, double-stranded DNA viruses that infect both vertebrate and invertebrate hosts (9, 64). The family Iridoviridae currently contains five genera, the Iridovirus and Chloriridovirus genera, associated with insects, the Lymphocystivirus and Megalocytivirus genera, which infect fish species, and the genus Ranavirus, whose members have been associated with mortality events in amphibians, fish, and reptiles (64). At this time, the type isolates for each genus in the family Iridoviridae have been sequenced (Table 1).

TABLE 1.

Completely sequenced iridoviruses

Genus Virus Known host Genome size (kb) GC content (%) No. of potential genes GenBank accession no.
Ranavirus ATV Salamander 106,332 54 92 AY150217
EHNV Fish 127,011 54 100 FJ433873
FV3 Frog 105,903 55 97 AY548484
TFV Frog 105,057 55 103 AF389451
SGIV Fish 140,131 48 139 AY521625
GIV Fish 139,793 49 139 AY666015
Megalocytivirus ISKNV Fish 111,362 55 117 AF371960
OSGIV Fish 112,636 54 116 AY894343
RBIV Fish 112,080 53 116 AY532606
Lymphocystivirus LCDV-1 Fish 102,653 29 108 L63545
LCDV-C Fish 186,247 27 178 AY380826
Iridovirus IIV-6 (CIV) Insect 212,482 29 211 AF303741
Chloriridovirus IIV-3 (MIV) Insect 190,132 48 126 DQ643392

Members of the genus Ranavirus have been recognized as major pathogens of economically and ecologically important cold-blooded vertebrates (8, 64). For example, ranaviruses (RVs) have been isolated from amphibians in North America (6, 18, 24, 34, 35), Asia (27, 66), Australia (56), and the United Kingdom (10, 19), from fish (2, 41, 46), and from reptiles (3, 14, 30, 37, 42, 43). In fact, ranaviruses are now considered agents of emerging infectious disease (9). As interest in RVs has grown, the number of ranaviruses that have been completely sequenced has also increased. These include frog virus 3 (FV3) (58), the type virus of the genus Ranavirus; tiger frog virus (TFV) (27), an RV closely related to FV3 that was isolated from frogs in Asia; and Ambystoma tigrinum virus (ATV) (36), an RV associated with salamander mortalities in North America. In addition, two grouper iridoviruses which are also members of the genus Ranavirus, the grouper iridovirus (GIV) (62) and the Singapore grouper iridovirus (SGIV) (53), were recently sequenced. In addition, at the time of preparation of the manuscript, the genomic sequence of the soft-shelled turtle ranavirus (STIV) became available (29). Information obtained by comparing ranavirus genomic sequences offers insight into RV evolutionary history, identifies core groups of genes, and gives insight into the genes responsible for viral immune evasion and pathogenesis.

Previous studies have shown that RV isolates can be translocated across large distances in infected salamanders that are used as bait for sport fishing (35, 44, 51). Phylogenetic analysis was used to compare the major capsid protein (MCP) sequences from salamander RV isolates from the southern Arizona border to Canada to other RV MCP sequences (35). The data suggest that salamander RV isolates are more closely related to fish RV isolates, such as epizootic hematopoietic necrosis virus (EHNV), than to other amphibian (frog) RV isolates, such as FV3 (35). Dot plot analysis comparing the genomic sequence of ATV to those of FV3 and TFV showed two major genomic inversions (36), while the FV3 and TFV genomes showed complete colinearity. These data suggest that at some point in virus evolutionary history, an ancestral virus diverged into the salamander virus and frog virus lineages. A genomic rearrangement occurred in one of the lineages at the time of divergence or after. Subsequent host-specific evolution occurred, limiting cross transmission among isolates, in such a way that frog RVs do not cause disease during laboratory infection of salamanders and vice versa (34). There is some evidence that salamander RV isolates can be isolated from or detected in laboratory-infected frogs (52) and that a pathogen host shift is the result of the movement of these pathogens (35). Thus, the ecological and economic consequences of RVs moving in the environment include the potential of these pathogens infecting and decimating new amphibian, fish, or reptile populations. Therefore, a more complete understanding of the genetic determinants that make up RVs would help to predict future transmission events.

EHNV was isolated in Australia from redfin perch (Perca fluviatilis) and rainbow trout (Oncorhynchus mykiss) (38, 39). EHNV can be classified as an indiscriminate pathogen of freshwater finfish, as it readily kills juvenile redfin perch and rainbow trout in inland water bodies throughout Australia (63). In addition, challenge experiments showed that following bath inoculation, other fish species are also susceptible to infection with EHNV, including the Macquarie perch (Macquaria australasica), silver perch (Bidyanus bidyanus), mosquito fish (Gambusia affinis), and mountain galaxias (Galaxias olidus). In contrast, Murray cod (Maccullochella peeli), golden perch (Macquaria ambigua), Australian bass (Macquaria novemaculeata), Macquarie perch, silver perch, and Atlantic salmon (Salmo salar) were susceptible only by intraperitoneal (i.p.) injection of virus. Serological surveys (A. Hyatt, unpublished data) show that redfin perch and rainbow trout can also be carriers of EHNV. Virus was reisolated from animals not showing clinical signs of disease, making them likely vehicles for the translocation and introduction of EHNV into naïve host populations. Preliminary and unpublished data have shown that i.p. inoculation of Australian frogs or the cane toad Bufo marinus with EHNV results in seroconversion but no signs of clinical disease (68). While EHNV has not been identified in fish populations in North America, it is possible that this pathogen could be translocated via movement of animals for food, bait, or scientific purposes, thereby infecting and potentially decimating naïve fish populations. In fact, the disease caused by EHNV is recognized by the World Organization for Animal Health (Office International Epizootics [OIE]) as a major cause of finfish mortalities (www.oie.int). In addition, both an EHNV disease in fish and a ranavirus infection in amphibians are notifiable diseases to OIE. Since the recognition of disease due to EHNV in Australia in 1986, similar systemic necrotizing iridovirus syndromes in farmed fish have been reported. These include catfish (Ictalurus melas) in France (European catfish virus) (45), sheatfish (Silurus glanis) in Germany (European sheatfish virus) (1), turbot (Scophthalmus maximus) in Denmark (5), and pike perch (Stizostedion lucioperca) in Finland (59). In addition, while EHNV has been classified as an RV, the relationship between this fish pathogen and amphibian RVs is poorly understood. Therefore, in order to better understand the relationships among RV isolates, the complete sequence of EHNV genomic DNA was determined. The characteristics of the EHNV genome, its relatedness to other iridoviruses, and insights into RV evolution are the focus of this study.

MATERIALS AND METHODS

Generation of EHNV genomic DNA library.

EHNV DNA was isolated as previously described (65) from cell culture-amplified virus stocks of the original EHNV isolated in Australia (39). The EHNV shotgun library was constructed by kinetically shearing 10 μg of viral DNA in 200 μl of TE (10 mM Tris-HCl, 1 mM EDTA) buffer. The sheared DNA was ethanol precipitated, and the pellet containing DNA was end repaired using T4 DNA polymerase and Klenow polymerase, concentrated by ethanol precipitation, and quantified. BstXI adaptors were ligated to the end-repaired viral DNA and then size selected by gel electrophoresis. DNAs of 2 to 4 kbp were extracted from the gel and ligated into the pOTWI3 plasmid. Plasmid DNA was transformed into DH10B competent cells by electroporation, plated on prewarmed agar plates containing 50 μg/ml chloramphenicol, and incubated overnight at 37°C. Colonies containing plasmid were selected using automated equipment, and plasmid DNA was isolated using solid-phase reversible immobilization (SPRI) technology. Isolated plasmids were sequenced from both ends of the insert by use of automated equipment (ABI 3730XL; Applied Biosystems). Sequences were aligned and assembled using Phred/Phrap (http://www.phrap.org) and finished using standard methods, with the aid of Consed (23).

Genome annotation.

The newly sequenced genome was annotated using similar procedures to those described previously (36). Using the BLASTP, BLASTX, and TBLASTX procedures (49, 50), all open reading frames (ORFs) with sequence similarity to any other closely related viral ORF and/or containing a domain(s) or homology with any known protein were identified. Identified ORFs were confirmed using the Genome Annotation Transfer Utility (GATU) (http://www.biovirus.org/), a program that uses previously annotated genomic DNA as a reference for annotating a newly sequenced genomic DNA, using all of the completely sequenced iridoviruses as reference sequences. The iridoviruses used in this analysis were as follows (also see Table 1): ATV (36), FV3 (58), TFV (27), GIV (62), SGIV (53), lymphocystis disease virus 1 (LCDV-1) (60), lymphocystis disease virus China (LCDV-C) (67), infectious spleen and kidney necrosis virus (ISKNV) (26), orange spotted grouper iridovirus (OSGIV) (40), rock bream iridovirus (RBIV) (15), insect iridovirus 6 (IIV-6) or Chilo iridovirus (CIV) (33), and invertebrate iridovirus 3 (IIV-3) or mosquito iridovirus (MIV) (13). The genome of a soft-shelled turtle ranavirus isolate was published during the preparation of the manuscript (29). However, due to the timing of this publication, the genomic information from this newly isolated RV was not used in our analysis. ORFs in the Iridoviridae family are presumed to be nonoverlapping; however, ORFs were considered overlapping if both ORFs had high sequence identity (i.e., a high BLASTP expect score) to other sequenced iridoviruses.

Phylogenetic and dot plot analysis.

Concatenated iridovirus phylogenetic analysis was conducted by obtaining the homologues of the EHNV ORFs 1L (myristylated membrane protein), 7R (RNA polymerase, α subunit), 8L (NTPase/helicase), 10L (DNA repair enzyme RAD2), 11R, 13L (ICP-46), 14L (MCP), 16L (thiol oxidoreductase), 18L (thymidine kinase), 19L (PCNA), 23L (transcription elongation factor SII), 24R (RNase III), 38R (ribonucleotide reductase, small subunit), 43R (RNA polymerase β subunit), 44L (DNA polymerase), 48L (CTD-phosphotransferase), 53L (myristylated membrane protein), 62R (tyrosine kinase), 72R, 77R, 85L (D5 NTPase), 86R, 89L (serine-threonine protein kinase), 92L (ABC ATPase), 95R, and 100R (putative replication factor) from representative members of the sequenced iridoviruses (see Tables S1 and S2 in the supplemental material) in GenBank by BLASTP analysis (http://www.ncbi.nlm.nih.gov/). All sequences were concatenated using BioEdit (http://www.mbio.ncsu.edu/BioEdit/bioedit.html). The sequences were aligned and neighbor-joining analysis was conducted using MEGA4 software (57) with default options.

Phylogenetic analysis of the EHNV ORF 87L was performed by acquiring homologous sequences by BLASTP analysis (http://www.ncbi.nlm.nih.gov/). The homologous sequences used in this phylogeny were from butterfish (GenBank accession no. ACQ58597.1), Atlantic salmon (NP_001135373.1), Northern pike (ACO14357.1), zebrafish (CAM14042.1), Norway rat (NP_569084.1), Chinese hamster (BAE78431.1), house mouse (NP_034179.1), rabbit (ACO49549.1), cow (P00376.3), horse (XP_001504693.1), rhesus monkey (XP_001110551.1), human (NP_000782.1), chimpanzee (XP_001134992.1), African clawed frog (NP_001088506.1), chicken (NP_001006584.2), zebra finch (XP_002190476.1), and herpesvirus saimiri 2 (NP_040203.1). The sequences were aligned and neighbor-joining analysis was conducted using MEGA4 software (57) with default options.

Dot plots using whole genome sequences comparing all of the sequenced iridoviruses (Table 1) to EHNV were generated using JDotter (http://www.biovirus.org/) (54, 55), using the default settings.

Nucleotide sequence accession number.

The EHNV sequence was deposited in the GenBank database (http://www.ncbi.nlm.nih.gov/) under accession number FJ433873.

RESULTS AND DISCUSSION

EHNV genome characteristics.

An EHNV random genomic DNA library was successfully generated and produced over 1,800 EHNV-specific sequences. The sequences contributed to the assembly of the complete genomic sequence, with an average final sequence coverage of >4-fold.

The finished EHNV genomic sequence (127,011 bp) is larger than the genomes of the amphibian RVs ATV, TFV, and FV3 (average, 105,754 bp), but smaller than the genomes of the grouper RVs GIV and SGIV (average, 139,962 bp) (Table 1). In addition, the EHNV genome is larger than the genomes of the other fish pathogens ISKNV, OSGIV, and RBIV (average, 112,026 bp) but smaller than the insect viral genomes of CIV and MIV (average, 201,307 bp). The lymphocystiviruses LCDV-1 and LCDV-C have very differently sized genomes (Table 1). EHNV genomic DNA is smaller than the average of these two viral genomic sequences (average, 144,450 bp). EHNV has a similar G+C content (54%) to those of TFV, FV3, ATV, ISKNV, OSGIV, and RBIV (53 to 55%), a slightly higher G+C content than those of the grouper iridoviruses GIV and SGIV and the insect iridovirus MIV (48 to 49%), and a much higher G+C content than those of LCDV-1, LCDV-C, and CIV (27 to 29%).

Open reading frame analysis.

One hundred ORFs are predicted for the EHNV genome, based on the annotation criteria described in Materials and Methods (Fig. 1; Table 2; see Table S1 in the supplemental material). The number of ORFs predicted for EHNV (100) is similar to the number of ORFs predicted for ATV (92) and close to the numbers of ORFs predicted for FV3 (97) and TFV (103) as well; however, all of these RVs have considerably fewer ORFs than the 139 ORFs seen in the fish RVs GIV and SGIV (Table 1). In addition, the number of EHNV ORFs is relatively close to the numbers of ORFs predicted for the other fish iridoviruses ISKNV, OSGIV, RBIV, and LCDV-1, while the numbers of ORFs in LCDV-C, CIV, and MIV are much larger, corresponding to their larger genome sizes (Table 1).

FIG. 1.

FIG. 1.

Annotation of the EHNV genome showing the order and orientation of ORFs. The EHNV genome was annotated as described in Materials and Methods. Each arrow represents an ORF in the right or left orientation. Conserved iridovirus ORFs (C) and putative virulence ORFs (V) are indicated. The brackets enclose the region of consecutively oriented ORFs and the arrows indicate the locations of genomic inversions compared to FV3.

TABLE 2.

Predicted EHNV open reading frames and best-matching iridovirus homologues

ORFa Position (bp) Size (no. of amino acids) MW Predicted function Best-matching iridovirus ORF Expect score % Identity % Positive Gap % GenBank accession no.
1L 11-1063 350 38,082 Myristylated membrane protein FV3 2L 1e−145 98 98 0 YP_031580
2L 1101-1940 279 31,302 FV3 2.5R 6e−163 98 99 0 AY548484
3R 1970-3184 404 44,679 TFV 4R 0.0 97 99 0 ABB92272
4R 3225-3407 60 6,615 ATV 4R 8e−18 95 95 0 YP_003775
5L 4077-4715 212 24,359 FV3 6R 4e−37 97 100 0 YP_031584
6L 5931-6353 140 14,881 ATV 5L 5e−48 92 94 0 YP_003776
7R 6433-10344 1,303 141,815 RNA polymerase α subunit ATV 6R 0.0 97 98 0 YP_003777
8L 11025-13871 948 106,432 NTPase or helicase ATV 7L 0.0 97 98 0 AAP33184
9R 13887-14300 137 15,018 TFV 10R 2e−70 99 99 0 ABB92276
10L 15273-16367 364 40,646 DNA repair enzyme RAD2 ATV 10L 0.0 97 99 0 AAP33187
11R 16461-16928 155 17,886 ATV 11R 7e−78 97 97 0 AAP33188
12R 17041-17208 55 5,686 TFV 99L 2e−09 96 98 0 ABB92345
13L 18518-19705 395 45,578 Immediate-early protein ICP-46 FV3 91R 0.0 97 98 0 AAT09751
14L 19829-21220 463 50,105 Major capsid protein ATV 14L 0.0 98 99 0 AAP33191
15L 21313-22419 368 41,742 TFV 95R 1e−135 79 80 14 ABB92343
16L 22487-22939 150 16,652 Thiol oxidoreductase ATV 16L 2e−83 99 100 0 AAP33193
17R 22972-24825 617 66,505 TFV 93L 0.0 96 97 0 ABB92341
18L 26068-26655 195 22,132 Deoxyribonucleoside kinase/thymidine kinase TFV 91.5R 7e−104 97 98 0 ABB92339
19L 26730-27512 260 27,685 Proliferating cell nuclear antigen (PCNA) ATV 20L 3e−133 96 98 0 AAP33197
20L 27933-28577 214 24,897 Cytosine DNA methyltransferase ATV 21L 9e−119 97 98 0 AAP33198
21L 29728-30621 297 33,872 Thymidylate synthase ATV 22L 4e−145 94 96 0 AAP33199
22L 30900-31373 157 17,405 Putative immediate-early protein ATV 23L 5e−89 99 100 0 AAP33200
23L 31502-31780 92 10,427 Transcription elongation factor SII ATV 24L 8e−41 98 98 0 AAP33201
24R 31836-32954 372 40,481 RNase III ATV 25R 0.0 98 99 0 AAP33202
25L 33681-35618 645 71,712 ATV 26L 0.0 82 85 10 AAP33203
26R* 36724-37491 255 29,090 TFV 83L 2e−118 94 95 0 ABB92336
27R 38298-38645 115 12,821 ATV 27R 1e−54 97 99 0 AAP33204
28L 38642-38863 73 7,952 ATV 28L 3e−35 98 98 0 AAP33205
29R 38926-39180 84 9,228 ATV 29R 8e−30 95 96 0 AAP33206
30R 39237-40418 393 42,021 ATV 30R 2e−164 92 95 0 AAP33207
31R 40638-41612 324 36,090 NTPase/helicase ATV 31R 1e−177 98 99 0 AAP33208
32R 42423-42938 171 18,773 FV3 72L 8e−78 97 99 0 AAT09732
33L 42429-42911 160 17,275 TFV 77L 2e−89 98 100 0 ABB92331
34L 43196-43432 78 8,381 ATV 34L 4e−24 92 94 0 AAP33211
35L 43473-43841 122 13,201 ATV 34.5L 9e−60 95 96 1 AY150217
36L 43859-44125 88 9,360 FV3 69R 7e−37 98 98 0 AAT09729
37R 44233-44937 234 25,229 ATV 37R 6e−112 90 95 0 AAP33214
38R 45326-46489 387 44,030 Ribonucleotide reductase, small subunit ATV 38R 0.0 98 99 0 AAP33216
39R 46547-47098 183 18,610 ATV38.5L 3e−49 96 97 0 AY150217
40R 47337-48059 240 22,522 FV3 65L 2e−07 97 97 0 AAT09725
41L 48535-48822 95 10,372 CARD-like caspase; putative interleukin-1 beta convertase ATV 40L 7e−42 94 97 0 YP_003812
42L 48917-49411 164 17,426 dUTPase FV3 63R 8e−79 96 98 0 YP_031642
43R 49791-53474 1227 133,829 DNA-dependent RNA polymerase, β subunit FV3 62L 0.0 97 98 0 YP_031641
44L 54549-57590 1013 114,508 DNA polymerase TFV 63R 0.0 99 99 0 AAL77804
45R 57756-58814 352 39,781 FV3 59L 0.0 97 98 0 YP_031638
46L 59343-59897 184 20,461 FV3 58.5R 2e−121 98 99 0 AY548484
47L 59912-60250 112 10,518 SGIV 45L 6e−11 51 59 0 YP_164140
48L 61123-62619 498 53,522 CTD-phosphotransferase ATV 47L 0.0 97 98 0 YP_003820
49L 62661-63065 134 15,244 TFV 58R 6e−66 97 98 0 ABB92316
50R 63102-63251 49 5,208 TFV 57L 4e−09 97 100 0 ABB92315
51R 63259-64554 431 47,190 Helicase ATV 50R 0.0 97 98 0 YP_003823
52R 64592-64822 76 8,757 Putative nuclear calmodulin-binding protein FV3 54L 1e−29 93 96 0 YP_031632
53L 65456-67027 523 54,869 Myristylated membrane protein ATV 51L 0.0 98 98 0 YP_003824
54R 67366-68433 355 39,399 3-β-Hydroxy-d-5-C27-steroid oxidoreductase-like protein FV3 52L 0.0 96 98 0 YP_031630
55R 68611-69672 353 39,507 FV3 23R 7e−46 32 50 7 YP_031601
56R 70259-71407 382 42,567 FV3 23R 0.0 97 97 0 YP_031601
57R 71804-72901 365 40,928 FV3 24R 0.0 96 99 0 YP_031602
58L 73215-73724 169 19,443 FV3 23R 2e−14 31 51 4 YP_031601
59L 74255-74989 244 27,214 TFV 23R 8e−33 34 56 2 ABB92288
60R 76159-77073 304 34,545 p31K ATV 55R 2e−148 96 98 0 YP_003828
61R 77399-78178 259 28,304 eIF2α homolog TFV 27R 4e−134 94 96 0 AAL77798
62R 78765-81677 970 106,902 Tyrosine kinase ATV 58R 0.0 96 98 0 YP_003831
63R 81726-82235 169 18,912 ATV 59R 8e−84 92 93 4 YP_003832
64R* 82345-83310 321 35,189 Putative capsid maturation protease (herpesvirus)
65R* 84057-84881 275 30,740
66R 84937-85422 161 17,937 SGIV 158L 2e−08 32 50 8 YP_164253
67R 86177-86596 139 15,143 ATV 60R 7e−73 98 99 0 YP_003833
68R 86646-88622 658 73,571 Neurofilament triplet H1-like protein TFV 33R 0.0 80 83 10 ABB92295
69R 88707-88898 63 6,613 TFV 34R 8e−12 95 96 0 ABB92296
70R 89046-89369 107 11,546 ATV 62.5R 9e−70 97 98 0 AY150217
71L 89432-89932 166 17,840 FV3 35L 7e−66 94 94 0 ABB92299
72R 90144-90779 211 23,478 TFV 40R 1e−109 98 99 0 ABB92302
73R 90918-92615 565 62,186 Ribonucleotide reductase, large subunit TFV 41R 0.0 98 99 0 AAL77800
74R 92722-93072 116 12,710 TFV 42R 6e−39 93 95 0 ABB92303
75R 93161-93967 268 28,876 ATV 67R 1e−62 76 79 5 YP_003841
76R 94224-94409 61 7,474 ATV 68R 2e−−08 72 72 24 YP_003842
77R 94481-97978 1165 129,034 ATV 69R 0.0 97 98 0 YP_003843
78L 98486-99343 285 29,429 TFV 46L 5e−74 69 70 21 ABB92307
79L 99469-99879 136 15,551 TFV 47L 2e−−72 99 100 0 ABB92308
80L 99933-100544 203 23,004 Neurofilament triplet H1-like protein ATV 72L 4e−36 65 70 26 YP_003846
81L 100670-101086 138 15,572 FV3 47L 1e−−60 97 99 0 YP_031625
82L 101089-101340 83 9,547 TFV 50L 9e−−35 96 98 0 ABB92310
83L 101459-103084 541 60,677 ATV 75L 5e−147 77 83 5 YP_003850
84R 103165-104850 561 61,508 ATV 76R 0.0 97 99 0 YP_003851
85L 105693-108614 973 108,742 D5 family NTPase FV3 22R 0.0 98 99 0 ABB92287
86R 108744-109403 219 25,327 ATV 78R 1e−104 99 99 0 YP_003853
87L* 110079-110648 189 21,084 Dihydrofolate reductase
88L 111114-111563 149 16,142 ATV 79L 1e−62 97 97 0 YP_003854
89L 111611-114334 907 98,760 Serine/threonine protein kinase TFV 19R 0.0 88 89 7 ABB92284
90R 114659-116167 502 53,345 ATV 81R 0.0 99 99 0 YP_003856
91R 116276-117151 291 31,665 ATV 82.5L 9e−147 96 98 0 YP_003857
92L 117540-118466 308 34,686 ABC-ATPase ATV 83L 3e−178 98 98 0 YP_003858
93L 118563-118919 118 13,354 FV3 14R 3e−44 97 98 0 YP_031592
94R 119046-119288 80 9,267 ATV 85.5L 4e−31 91 96 0 YP_003860
95R 119947-120840 297 32,649 ATV 87R 2e−132 96 98 0 YP_003862
96L 120906-121118 70 7,885 ATV 88L 9e−24 95 98 0 YP_003863
97L 121917-122354 145 16,101 LCDV-C 71L 0.22 29 52 6 YP_073578
98R 123028-123714 228 24,474 ATV 89R 8e−121 96 97 0 YP_003864
99R 123780-124193 137 15,166 ATV 90R 3e−70 94 97 0 YP_003865
100R 124882-125652 256 29,693 Putative replication factor FV3 1R 5e−144 98 98 0 YP_031579
a

*, EHNV-specific ORFs.

Of the 100 EHNV ORFs, 26 ORFs are conserved throughout the family Iridoviridae (see Tables S2 and S3 in the supplemental material). These ORFs can be defined as the core iridovirus genes, since they are present in every iridovirus sequenced to date and confirm published reports that all iridoviruses contain these 26 ORFs (20). The majority of these conserved ORFs (21/26 ORFs) have a predicted function, based on sequence homology to other characterized proteins, or have been identified based on experimental data (see Table S2 in the supplemental material). In contrast, only 4 of the 27 additional ORFs that are conserved throughout the genus Ranavirus have a predicted function (see Tables S1 and S4 in the supplemental material), and only 1 of the 13 amphibian RV-specific genes has a predicted function (the viral homologue of eukaryotic translation initiation factor 2α [vIF2αH; ORF 61R]) (see Table S5 in the supplemental material).

Interestingly, there are three unique EHNV ORFs, namely, ORFs 64R and 65R, which have no homology to any known protein sequence, and ORF 87L, which is predicted to encode a dihydrofolate reductase (DHFR) (Table 2; see Table S1 in the supplemental material). DHFRs are thought to be involved in nucleotide metabolism and have been described for herpesviruses (61). BLAST search analysis of the EHNV ORF 87L showed the highest sequence similarity with butterfish, Atlantic salmon, and northern pike DHFRs (data not shown). Phylogenetic analysis of this EHNV ORF compared to homologous sequences suggests that EHNV acquired this unique gene from a fish host, thereby allowing the virus to replicate and cause disease in finfish (Fig. 2). There is evidence of host-derived gene transfer in poxviruses (7), a group of closely related DNA viruses (31, 32), so it is reasonable to hypothesize that similar events have occurred with iridoviruses. In fact, iridoviruses may have a higher rate of horizontal gene transfer from their host due to the nuclear stage of iridovirus DNA replication (9, 64).

FIG. 2.

FIG. 2.

Phylogenetic analysis of EHNV ORF 87L. Homologous sequences to the EHNV DHFR gene (ORF 87L) were obtained by BLASTP analysis. The neighbor-joining tree was determined using MEGA4, and it is shown with statistical support indicating the robustness of the inferred branching pattern, as assessed using the bootstrap test. The accession number for each gene in the phylogeny is given in Materials and Methods.

There appear to be nine ORF clusters, containing a minimum of four consecutively oriented ORFs (COOs), throughout the genome. These clusters of COOs all have similar orientations, either right or left, with the majority of the COOs having the same orientation. While the genomes of iridoviruses are circularly permuted and terminally redundant (64), and therefore the orientation of these ORFs relative to the orientation of the start of the genome was an arbitrary decision, the amount of conservation among RV isolates within these regions is surprising. For example, the region between EHNV ORFs 54 and 77 contains 21 of the 24 predicted ORFs in the right orientation, while in ATV 19 of the 20 ORFs in this region (ATV ORFs 52 to 69) are oriented in the same direction (36). FV3 and TFV also have similarly oriented ORFs in this region (27, 58).

In overall appearance, the EHNV COOs are reminiscent of pathogenesis islands (PAIs) found in pathogenic bacteria. The bacterial PAIs, mobile genetic elements that contribute to rapid changes in virulence potential (16, 21, 22, 25), contain ORFs that are in the same orientation and code for proteins that have been correlated with increased pathogenesis. Poxviruses contain groups of ORFs at the hairpin ends of their genomes that are associated with virulence and have been suggested to be analogous to PAIs (11). It is interesting to note the similarity between these related vertebrate pathogens in that ORFs correlating with pathogenesis are in close proximity to each other. Further analysis of this region may help to shed light on the specific function of these ORFs.

Another interesting region of the EHNV genome is between ORFs 55R and 59L. The five ORFs within this region share homology with two ORFs from ATV (53R and 54R) and two ORFs from FV3 and TFV (23R and 24R), as well as sharing homology with each other (see Table S1 in the supplemental material). In addition, these EHNV ORFs also share homology with SGIV and GIV ORFs. EHNV 57R has homology with GIV 53L, while EHNV 56R has homology to five GIV ORFs (GIV ORFs 50L, 51L, 53L, 91L, and 94L) (see Table S1 in the supplemental material). Alignments of these EHNV ORFs (data not shown) suggest that these five EHNV ORFs may be the result of gene duplication events.

Phylogenetic analysis.

Twenty-six EHNV ORFs, the core iridovirus genes, and the orthologous genes from the 12 completely sequenced iridoviruses were used to generate a concatenated phylogeny (Fig. 3). The phylogenetic tree shows high bootstrap support (100%) for EHNV being a member of the genus Ranavirus in the family Iridoviridae. In addition, EHNV is more closely related to ATV, a salamander RV, than it is to FV3 and TFV, which are frog RVs, supporting previously published analysis (35). EHNV is more distantly related to LCDV-1 and LVDC-C, to ISKNV, to OSGIV, and to RBIV, MIV, and CIV (members of the Lymphocystivirus, Megalocytivirus, Chloriridovirus, and Iridovirus genera, respectively). The grouper iridoviruses GIV and SGIV are clearly more divergent from the ATV/EHNV and FV3/TFV ranaviruses. Based on these data and that of others (53, 62), the grouper iridoviruses or GIV-like isolates could be considered a distantly related species within the genus Ranavirus, or perhaps a subspecies of ranaviruses. Therefore, we suggest that the RVs be divided into two subspecies, the GIV-like RV and the amphibian-like ranavirus (ALRV) subspecies, based on our observations and those of others (20). The ALRVs can then be classified further as being ATV-like or FV3-like. It is interesting that the branch lengths of the ALRVs are very short, suggesting evolutionarily recent speciation of these viral isolates.

FIG. 3.

FIG. 3.

Concatenated phylogeny of 26 conserved iridovirus sequences. Phylogenetic relationships of 26 conserved open reading frames from the 13 completely sequenced iridovirus genomes are shown. The neighbor-joining tree obtained using MEGA4 is shown, with statistical support indicating the robustness of the inferred branching pattern, as assessed using the bootstrap test. The sequences used for this analysis are described in Tables S1 and S3 in the supplemental material. The most recent common ancestors (MRCAs) are indicated at particular branch points on the phylogeny.

Whole-genome alignments.

Dot plot comparisons of whole genomic sequences can reveal a large amount of information on the entire genome and how genomic sequences are organized (e.g., colinearity, inversions, and repeat sequences). Comparing the sequence of EHNV to that of ATV, a −45° line can be observed (Fig. 4). Breaks in this line reveal inserted sequences, i.e., ORFs present in one genome and not in the other (Fig. 4; see Table S1 in the supplemental material). In addition, repeated “dots” can be seen running vertically and horizontally throughout the dot plot (Fig. 4). Interestingly, one can almost map out the EHNV ORFs between these “dots,” suggesting that these repeated regions may be involved in the regulation of gene expression (Fig. 4). The “dots” also appear to be running in the lighter streaks running vertically and horizontally in the dot plot (Fig. 4). These lighter streaks represent regions of different G/C or A/T content, a phenomenon also observed with poxviruses (11). The EHNV genome has an overall G+C content of 54% (Table 1), but this does not mean that the entire genome has a uniform %GC, as some regions are more G/C rich than others. The A/T-rich regions appear as lighter streaks on the dot plot, with the off-vertical dots indicating similar A/T-rich regions at multiple places in the genomes. Similar patterns have been observed previously in dot plots for ATV, TFV, and FV3 (20, 36, 58).

FIG. 4.

FIG. 4.

Dot plot analysis of EHNV versus ATV. The genomic sequence of EHNV was compared to that of ATV by dot plot analysis (JDotter [www.biovirus.org/]). The dot plot comparison of EHNV with ATV shows unique sequences in the EHNV genome.

As observed in other dot plot comparisons (20, 36), the typical RV repeat patterns can be observed by comparing EHNV to ATV, FV3, and TFV (Fig. 4 and 5). Comparing EHNV to ATV by dot plot analysis showed colinearity between these two RVs (Fig. 4). This is a surprising result, as these two viruses infect very different hosts and have been isolated on different continents. This observation suggests that these two different RV pathogens are very closely related, confirming the phylogenetic analysis of 26 ORFs (Fig. 3). There are regions of the dot plot that show unique sequences in EHNV compared to ATV, and these regions correlate with the unique EHNV ORFs (Fig. 4; see Table S1 in the supplemental material) or with extra noncoding DNA sequences. These extra sequences are visualized as breaks in the −45° colinear line and a shift in this line to the right. This shift represents a sequence that is in EHNV but not present in ATV (Fig. 4). There are no sequences present in ATV that are missing from EHNV.

FIG. 5.

FIG. 5.

Dot plot analysis of EHNV versus other ranaviruses. The genomic sequence of EHNV was compared to those of FV3 and SGIV by dot plot analysis (JDotter [www.biovirus.org/]). (A) Comparison of the EHNV genome to the FV3 genome. (B) Dot plot comparison of EHNV and SGIV genomic sequences.

In contrast to previous reports (58), no inversions were observed between FV3 and TFV, using the same program (MacVector) or the dot plot program used in this study (JDotter; data not shown). In addition, dot plots comparing FV3 and STIV revealed colinearity between these two ranaviruses (29), and dot plots comparing SGIV with GIV also revealed complete colinearity, although the starts of these RV genomes differ (data not shown). Therefore, EHNV and ATV, FV3, STIV, and TFV, and SGIV and GIV were each grouped together for dot plot analysis (Fig. 5A and B). Two major genomic inversions in the FV3/TFV lineage compared to the EHNV/ATV lineage can be visualized on the dot plot as a +45° line (Fig. 5A). This is similar to previously published reports comparing FV3/TFV with ATV (20, 36). In comparing EHNV/ATV to GIV/SGIV, long stretches of colinearity are not observed between these sequences, although small sections of colinearity can be observed in the dot plot (Fig. 5B). The short diagonal lines on the dot plot are indicative of groups of ORFs (2 to 4 ORFs) that are scattered throughout the genome, suggesting that major genomic rearrangements have taken place among RV species. The dot plot correlates with the phylogeny in that EHNV is more closely related to the amphibian RVs than it is to the GIV-like viruses that infect fish.

Very little colinearity was observed in comparing EHNV to all other completely sequenced iridovirus isolates (data not shown). Short stretches of colinearity were observed, but the numbers, intensities, and lengths of these lines were much smaller when comparing EHNV to all other iridovirus genomic sequences.

A closer examination of the genomic dot plots between SGIV and EHNV and between SGIV and FV3 revealed additional information on the evolution of the ranaviruses. In Fig. 6A and B, the small stretches of colinear ORFs along the SGIV genome are numbered consecutively. EHNV (and ATV) has segments 1 through 6 oriented together in consecutive order (Fig. 6A). However, the FV3 genome has a rearrangement of these segments (Fig. 6B). This rearrangement of segments corresponds to the inversion observed when comparing the EHNV/ATV and FV3/TFV genomic sequences (Fig. 5A and arrows in Fig. 6). Therefore, these data suggest that in EHNV/ATV, the gene order in this region is similar to the gene order in the most common recent ancestor (MRCA) of the ranaviruses (MRCA A) (Fig. 3) and that the inversion observed when comparing EHNV/ATV and FV3/TFV occurred in the FV3-like lineage (Fig. 7). There is not enough conservation of gene order in the region of the second inversion to be able to establish if the second inversion occurred in the ATV-like or FV3-like lineage.

FIG. 6.

FIG. 6.

Dot plot analysis of SGIV compared to EHNV and FV3. Dot plots were generated comparing SGIV to EHNV (A) and FV3 (B). Colinear segments were sequentially numbered along the SGIV genome. Consecutively ordered segments along the EHNV and FV3 genomes are circled, while inverted segments are boxed.

FIG. 7.

FIG. 7.

Model of ranavirus genomic rearrangements. Using the dot plot analysis shown in Fig. 5, consecutively ordered colinear segments were arranged diagrammatically. Comparing the order and orientation of the colinear segments, the rearrangements observed in Fig. 4 occurred in the FV3-like virus lineage and not in the ATV-like virus lineage.

Since the ALRVs contain a virus isolated from fish (EHNV) and since all of the more distantly related vertebrate iridoviruses infect fish (GIV/SGIV, LCDV, and ISKNV), we hypothesize that the most recent common ancestor of the ALRVs was an ancestral fish virus (MRCA B) (Fig. 3 and 8A). Thus, an evolutionarily recent host shift from fish to amphibians, surmised by the shallowness of the ALRV branch lengths, must have occurred. In fact, our data suggest that there were two species jumps, one from fish to frogs (MRCA B) and a jump from fish to salamanders (MRCA C) (Fig. 8A). In addition, the rearrangement of the FV3-like virus genomic DNA relative to ATV/EHNV resulted in the speciation of the ALRVs into the ATV-like and FV3-like virus lineages. The newly acquired sequence of the soft-shelled turtle RV, which is completely colinear with FV3 (29), suggests that another host jump, from frogs to reptiles, also took place recently in evolutionary history (Fig. 3 and 8). However, an alternative hypothesis, where the most recent common ancestor for the ALRVs (MRCA B in Fig. 3 and 8B) infected a tetrapod amphibian, such that the species jump from fish to amphibians occurred prior to speciation of the ALRVs, is also consistent with the phylogeny in Fig. 3. This alternative hypothesis would require a more recent jump back from tetrapod amphibians into fish, yielding an EHNV-like virus. Both of these hypotheses suggest that for the majority of evolutionary time vertebrate iridoviruses were confined to fish, and much more recently, there appear to have been at least three species jumps, from fish to frogs, from fish to salamanders, and from frogs to reptiles, and perhaps as many as four species jumps, including a jump from tetrapod amphibians back to fish. It is tempting to speculate that activities associated with human harvesting of aquatic organisms during the past 40,000 years (28, 47, 48) led to the more common recent jumping of ranaviruses among aquatic organisms.

FIG. 8.

FIG. 8.

Ranavirus multiple-species jump hypotheses. Throughout the majority of evolutionary history, the iridoviruses have been restricted to fish species. Based on recent genomic sequence information, we hypothesize that the most recent common ancestor of the ranaviruses was a fish virus (MRCA A). In addition, we hypothesize that there have been evolutionarily recent host shifts. We propose two hypotheses to explain these multiple recent host shifts among the amphibian-like ranaviruses. (A) One hypothesis suggests that the most recent common ancestor of the ALRVs was a fish virus (MRCA B) and that a jump occurred from fish to frogs, with a subsequent jump from frogs to turtles. In addition, if the most recent common ancestor of the ATV-like viruses was a fish virus (MRCA C), then another jump from fish to salamanders occurred. (B) An alternative hypothesis suggests that the most recent common ancestor of the ranaviruses was a fish virus (MRCA A) and that a jump occurred from fish into tetrapod amphibians (MRCA B). At this time, it is unclear if the shift in host species was from fish to frogs, fish to salamanders, or both. A subsequent host shift occurred from frogs to turtles, as well as a jump from salamanders back into fish.

The sequencing of EHNV has allowed us to hypothesize that ranavirus host shifts are possible and that there have been evolutionarily recent ranavirus host shifts. In addition, the ability of this group of viruses to infect such a wide variety of host species suggests that more host shifts are likely. Therefore, it is important that we understand more of the evolutionary traits of this unique group of viruses, as there is no other closely related group of viruses that infect such a broad group of hosts, with the possible exception of the orthomyxoviruses (64).

Infectious diseases have become recognized as one of the most important threats to public, veterinary, and wildlife health over the past 30 years (4, 12). Combating infectious diseases is a key goal of public and veterinary health efforts, both nationally and internationally. Infectious diseases in insects, mammals, marsupials, amphibians, reptiles, and fish are not well understood. Global air travel, trade, tourism, immigration, and expansion of human settlements effectively increase the mixing of pathogens among humans as well as domestic and wild animals. From 1998 to 2000, the most-reported wildlife pathogens were viruses related to the anthropogenic movement of animals (12, 17). Recent reports have suggested that RVs move around the globe in host species used for bait, food, pets, and research (35, 44, 51). This phenomenon may increase the probability of new RV pathogens emerging in naïve populations and supports the need for controlling the movement of RV host species within and between geographical regions. Since RVs infect a wide variety of ecologically and economically important hosts, understanding RV evolution, including the importance of the unique genomic rearrangements found among RV isolates in relation to host specificity and viral evolution, will help to predict and perhaps to prevent further RV epizootics. While this study does give insight into RV evolution, more genomic sequence information is needed to continue our efforts to understand the role that RVs play in the environment.

Supplementary Material

[Supplemental material]

Acknowledgments

This work was supported in part by Integrated Research Challenges in Environmental Biology (IBN-9977063) and Division of Environmental Biology (0213851) grants from the National Science Foundation.

Footnotes

Published ahead of print on 30 December 2009.

Supplemental material for this article may be found at http://jvi.asm.org/.

REFERENCES

  • 1.Ahne, W., M. Bearzotti, M. Bremont, and S. Essbauer. 1989. Comparison of European systemic piscine and amphibian iridoviruses with epizootic haematopoietic necrosis virus and frog virus 3. Zentralbl. Veterinarmed. B 45:373-383. [DOI] [PubMed] [Google Scholar]
  • 2.Ahne, W., M. Bremont, R. P. Hedrick, A. D. Hyatt, and R. J. Whittington. 1997. Iridoviruses associated with epizootic haematopoietic necrosis (EHN) in aquaculture. World J. Microbiol. Biotechnol. 13:367-373. [Google Scholar]
  • 3.Allender, M. C., M. M. Fry, A. R. Irizarry, L. Craig, A. J. Johnson, and M. Jones. 2006. Intracytoplasmic inclusions in circulating leukocytes from an eastern box turtle (Terrapene carolina carolina) with iridoviral infection. J. Wildl. Dis. 42:677-684. [DOI] [PubMed] [Google Scholar]
  • 4.Binder, S., A. M. Levitt, J. J. Sacks, and J. M. Hughes. 1999. Emerging infectious diseases: public health issues for the 21st century. Science 284:1311-1313. [DOI] [PubMed] [Google Scholar]
  • 5.Bloch, B., and J. L. Larsen. 1993. An iridovirus-like agent associated with systemic infection in cultured turbot Scophthalmus-maximus fry in Denmark. Dis. Aquat. Organ. 15:235-240. [Google Scholar]
  • 6.Bollinger, T. K., J. Mao, D. Schock, R. M. Brigham, and V. G. Chinchar. 1999. Pathology, isolation, and preliminary molecular characterization of a novel iridovirus from tiger salamanders in Saskatchewan. J. Wildl. Dis. 35:413-429. [DOI] [PubMed] [Google Scholar]
  • 7.Bratke, K. A., and A. McLysaght. 2008. Identification of multiple independent horizontal gene transfers into poxviruses using a comparative genomics approach. BMC Evol. Biol. 8:67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Chinchar, V. G. 2002. Ranaviruses (family Iridoviridae): emerging cold-blooded killers. Arch. Virol. 147:447-470. [DOI] [PubMed] [Google Scholar]
  • 9.Chinchar, V. G., A. Hyatt, T. Miyazaki, and T. Williams. 2009. Family Iridoviridae: poor viral relations no longer. Curr. Top. Microbiol. Immunol. 328:123-170. [DOI] [PubMed] [Google Scholar]
  • 10.Cunningham, A. A., T. E. S. Langton, P. M. Bennett, J. F. Lewin, S. E. N. Drury, R. E. Gough, and S. K. MacGregor. 1996. Pathological and microbiological findings from incidents of unusual mortality of the common frog (Rana temporaria). Philos. Trans. R. Soc. Lond. B 351:1539-1557. [DOI] [PubMed] [Google Scholar]
  • 11.Da Silva, M., and C. Upton. 2005. Host-derived pathogenicity islands in poxviruses. Virol. J. 2:30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Daszak, P., A. A. Cunningham, and A. D. Hyatt. 2000. Emerging infectious diseases of wildlife—threats to biodiversity and human health. Science 287:443-449. [DOI] [PubMed] [Google Scholar]
  • 13.Delhon, G., E. R. Tulman, C. L. Afonso, Z. Lu, J. J. Becnel, B. A. Moser, G. F. Kutish, and D. L. Rock. 2006. Genome of invertebrate iridescent virus type 3 (mosquito iridescent virus). J. Virol. 80:8439-8449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.De Voe, R., K. Geissler, S. Elmore, D. Rotstein, G. Lewbart, and J. Guy. 2004. Ranavirus-associated morbidity and mortality in a group of captive eastern box turtles (Terrapene carolina carolina). J. Zoo Wildl. Med. 35:534-543. [DOI] [PubMed] [Google Scholar]
  • 15.Do, J. W., C. H. Moon, H. J. Kim, M. S. Ko, S. B. Kim, J. H. Son, J. S. Kim, E. J. An, M. K. Kim, S. K. Lee, M. S. Han, S. J. Cha, M. S. Park, M. A. Park, Y. C. Kim, J. W. Kim, and J. W. Park. 2004. Complete genomic DNA sequence of rock bream iridovirus. Virology 325:351-363. [DOI] [PubMed] [Google Scholar]
  • 16.Dobrindt, U., B. Hochhut, U. Hentschel, and J. Hacker. 2004. Genomic islands in pathogenic and environmental microorganisms. Nat. Rev. Microbiol. 2:414-424. [DOI] [PubMed] [Google Scholar]
  • 17.Dobson, A., and J. Foufopoulos. 2001. Emerging infectious pathogens of wildlife. Philos. Trans. R. Soc. Lond. B 356:1001-1012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Donnelly, T. M., E. W. Davidson, J. K. Jancovich, S. Borland, M. Newberry, and J. Gresens. 2003. What's your diagnosis? Ranavirus infection. Lab. Anim. (New York) 32:23-25. [DOI] [PubMed] [Google Scholar]
  • 19.Drury, S. E. N., R. E. Gough, and A. A. Cunningham. 1995. Isolation of an iridovirus-like agent from common frogs (Rana-temporaria). Vet. Rec. 137:72-73. [DOI] [PubMed] [Google Scholar]
  • 20.Eaton, H. E., J. Metcalf, E. Penny, V. Tcherepanov, C. Upton, and C. R. Brunetti. 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]
  • 21.Gerlach, R. G., and M. Hensel. 2007. Salmonella pathogenicity islands in host specificity host pathogen-interactions and antibiotic resistance of Salmonella enterica. Berl. Munch. Tierarztl. Wochenschr. 120:317-327. [PubMed] [Google Scholar]
  • 22.Gerlach, R. G., D. Jackel, B. Stecher, C. Wagner, A. Lupas, W. D. Hardt, and M. Hensel. 2007. Salmonella pathogenicity island 4 encodes a giant non-fimbrial adhesin and the cognate type 1 secretion system. Cell. Microbiol. 9:1834-1850. [DOI] [PubMed] [Google Scholar]
  • 23.Gordon, D., C. Abajian, and P. Green. 1998. Consed: a graphical tool for sequence finishing. Genome Res. 8:195-202. [DOI] [PubMed] [Google Scholar]
  • 24.Greer, A. L., M. Berrill, and P. J. Wilson. 2005. Five amphibian mortality events associated with ranavirus infection in south central Ontario, Canada. Dis. Aquat. Organ. 67:9-14. [DOI] [PubMed] [Google Scholar]
  • 25.Hacker, J., B. Hochhut, B. Middendorf, G. Schneider, C. Buchrieser, G. Gottschalk, and U. Dobrindt. 2004. Pathogenomics of mobile genetic elements of toxigenic bacteria. Int. J. Med. Microbiol. 293:453-461. [DOI] [PubMed] [Google Scholar]
  • 26.He, J. G., M. Deng, S. P. Weng, Z. Li, S. Y. Zhou, Q. X. Long, X. Z. Wang, and S. M. Chan. 2001. Complete genome analysis of the mandarin fish infectious spleen and kidney necrosis iridovirus. Virology 291:126-139. [DOI] [PubMed] [Google Scholar]
  • 27.He, J. G., L. Lu, M. Deng, H. H. He, S. P. Weng, X. H. Wang, S. Y. Zhou, Q. X. Long, X. Z. Wang, and S. M. Chan. 2002. Sequence analysis of the complete genome of an iridovirus isolated from the tiger frog. Virology 292:185-197. [DOI] [PubMed] [Google Scholar]
  • 28.Hu, Y. W., H. Shang, H. W. Tong, O. Nehlich, W. Liu, C. H. Zhao, J. C. Yu, C. S. Wang, E. Trinkaus, and M. P. Richards. 2009. Stable isotope dietary analysis of the Tianyuan 1 early modern human. Proc. Natl. Acad. Sci. USA 106:10971-10974. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Huang, Y. H., X. H. Huang, H. Liu, J. Gong, Z. L. Ouyang, H. C. Cui, J. H. Cao, Y. T. Zhao, X. J. Wang, Y. L. Jiang, and Q. W. Qin. 2009. Complete sequence determination of a novel reptile iridovirus isolated from soft-shelled turtle and evolutionary analysis of Iridoviridae. BMC Genomics 10:224-238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Hyatt, A. D., M. Williamson, B. E. H. Coupar, D. Middleton, S. G. Hengstberger, A. R. Gould, P. Selleck, T. G. Wise, J. Kattenbelt, A. A. Cunningham, and J. Lee. 2002. First identification of a ranavirus from green pythons (Chondropython viridis). J. Wildl. Dis. 38:239-252. [DOI] [PubMed] [Google Scholar]
  • 31.Iyer, L. A., S. Balaji, E. V. Koonin, and L. Aravind. 2006. Evolutionary genomics of nucleo-cytoplasmic large DNA viruses. Virus Res. 117:156-184. [DOI] [PubMed] [Google Scholar]
  • 32.Iyer, L. M., L. Aravind, and E. V. Koonin. 2001. Common origin of four diverse families of large eukaryotic DNA viruses. J. Virol. 75:11720-11734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Jakob, N. J., K. Muller, U. Bahr, and G. Darai. 2001. Analysis of the first complete DNA sequence of an invertebrate iridovirus: coding strategy of the genome of Chilo iridescent virus. Virology 286:182-196. [DOI] [PubMed] [Google Scholar]
  • 34.Jancovich, J. K., E. W. Davidson, J. F. Morado, B. L. Jacobs, and J. P. Collins. 1997. Isolation of a lethal virus from the endangered tiger salamander Ambystoma tigrinum stebbinsi. Dis. Aquat. Organ. 31:161-167. [Google Scholar]
  • 35.Jancovich, J. K., E. W. Davidson, N. Parameswaran, J. Mao, V. G. Chinchar, J. P. Collins, B. L. Jacobs, and A. Storfer. 2005. Evidence for emergence of an amphibian iridoviral disease because of human-enhanced spread. Mol. Ecol. 14:213-224. [DOI] [PubMed] [Google Scholar]
  • 36.Jancovich, J. K., J. Mao, V. G. Chinchar, C. Wyatt, S. T. Case, S. Kumar, G. Valente, S. Subramanian, E. W. Davidson, J. P. Collins, and B. L. Jacobs. 2003. Genomic sequence of a ranavirus (family Iridoviridae) associated with salamander mortalities in North America. Virology 316:90-103. [DOI] [PubMed] [Google Scholar]
  • 37.Johnson, A. J., A. P. Pessier, and E. R. Jacobson. 2007. Experimental transmission and induction of ranaviral disease in western ornate box turtles (Terrapene ornata ornata) and red-eared sliders (Trachemys scripta elegans). Vet. Pathol. 44:285-297. [DOI] [PubMed] [Google Scholar]
  • 38.Langdon, J. S., J. D. Humphrey, and L. M. Williams. 1988. Outbreaks of an EHNV-like iridovirus in cultured rainbow-trout, Salmo-gairdneri richardson, in Australia. J. Fish Dis. 11:93-96. [Google Scholar]
  • 39.Langdon, J. S., J. D. Humphrey, L. M. Williams, A. D. Hyatt, and H. A. Westbury. 1986. 1st virus isolation from Australian fish—an iridovirus-like pathogen from redfin perch, Perca-fluviatilis L. J. Fish Dis. 9:263-268. [Google Scholar]
  • 40.Lu, L., S. Y. Zhou, C. Chen, S. P. Weng, S. M. Chan, and J. G. He. 2005. Complete genome sequence analysis of an iridovirus isolated from the orange-spotted grouper, Epinephelus coioides. Virology 339:81-100. [DOI] [PubMed] [Google Scholar]
  • 41.Mao, J. H., R. P. Hedrick, and V. G. Chinchar. 1997. Molecular characterization, sequence analysis, and taxonomic position of newly isolated fish iridoviruses. Virology 229:212-220. [DOI] [PubMed] [Google Scholar]
  • 42.Marschang, R. E., P. Becher, H. Posthaus, P. Wild, H. J. Thiel, U. Muller-Doblies, E. F. Kalet, and L. N. Bacciarini. 1999. Isolation and characterization of an iridovirus from Hermann's tortoises (Testudo hermanni). Arch. Virol. 144:1909-1922. [DOI] [PubMed] [Google Scholar]
  • 43.Marschang, R. E., S. Braun, and P. Becher. 2005. Isolation of a ranavirus from a gecko (Uroplatus fimbriatus). J. Zoo Wildl. Med. 36:295-300. [DOI] [PubMed] [Google Scholar]
  • 44.Picco, A. M., and J. P. Collins. 2008. Amphibian commerce as a likely source of pathogen pollution. Conserv. Biol. 22:1582-1589. [DOI] [PubMed] [Google Scholar]
  • 45.Pozet, F., M. Morand, A. Moussa, C. Torhy, and P. Dekinkelin. 1992. Isolation and preliminary characterization of a pathogenic icosahedral deoxyribovirus from the catfish Ictalurus-melas. Dis. Aquat. Organ. 14:35-42. [Google Scholar]
  • 46.Qin, Q. W., S. F. Chang, G. H. Ngoh-Lim, S. Gibson-Kueh, C. Shi, and T. J. Lam. 2003. Characterization of a novel ranavirus isolated from grouper Epinephelus tauvina. Dis. Aquat. Organ. 53:1-9. [DOI] [PubMed] [Google Scholar]
  • 47.Richards, M. P., P. B. Pettitt, M. C. Stiner, and E. Trinkaus. 2001. Stable isotope evidence for increasing dietary breadth in the European mid-Upper Paleolithic. Proc. Natl. Acad. Sci. USA 98:6528-6532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Richards, M. P., and E. Trinkaus. 2009. Isotopic evidence for the diets of European Neanderthals and early modern humans. Proc. Natl. Acad. Sci. USA 106:16034-16039. doi: 10.1073/pnas.0903821106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Schaffer, A. A., L. Aravind, T. L. Madden, S. Shavirin, J. L. Spouge, Y. I. Wolf, E. V. Koonin, and S. F. Altschul. 2001. Improving the accuracy of PSI-BLAST protein database searches with composition-based statistics and other refinements. Nucleic Acids Res. 29:2994-3005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Schaffer, A. A., Y. I. Wolf, C. P. Ponting, E. V. Koonin, L. Aravind, and S. F. Altschul. 1999. IMPALA: matching a protein sequence against a collection of PSI-BLAST-constructed position-specific score matrices. Bioinformatics 15:1000-1011. [DOI] [PubMed] [Google Scholar]
  • 51.Schloegal, L. M., A. M. Picco, A. M. Kilpatrick, A. J. Davies, and A. D. Hyatt. 2009. Magnitude of the US trade in amphibians and presence of Batrachochytrium dendrobatidis and ranavirus infection in imported North American bullfrogs (Rana catesbeiana). Biol. Conserv. 142:1420-1426. [Google Scholar]
  • 52.Schock, D. M., T. K. Bollinger, V. G. Chinchar, J. K. Jancovich, and J. P. Collins. 2008. Experimental evidence that amphibian ranaviruses are multi-host pathogens. Copeia 2008:133-143. [Google Scholar]
  • 53.Song, W. J., Q. W. Qin, J. Qiu, C. H. Huang, F. Wang, and C. L. Hew. 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]
  • 54.Sonnhammer, E. L. L., and R. Durbin. 1995. A dot-matrix program with dynamic threshold control suited for genomic DNA and protein-sequence analysis. Gene 167:GC1-GC10. [DOI] [PubMed] [Google Scholar]
  • 55.Sonnhammer, E. L. L., and J. C. Wootton. 2001. Integrated graphical analysis of protein sequence features predicted from sequence composition. Proteins 45:262-273. [DOI] [PubMed] [Google Scholar]
  • 56.Speare, R., and J. R. Smith. 1992. An iridovirus-like agent isolated from the ornate burrowing frog Limnodynastes-ornatus in Northern Australia. Dis. Aquat. Organ. 14:51-57. [Google Scholar]
  • 57.Tamura, K., J. Dudley, M. Nei, and S. Kumar. 2007. MEGA4: molecular evolutionary genetics analysis (MEGA) software version 4.0. Mol. Biol. Evol. 24:1596-1599. [DOI] [PubMed] [Google Scholar]
  • 58.Tan, W. G. H., T. J. Barkman, V. G. Chinchar, and K. Essani. 2004. Comparative genomic analyses of frog virus 3, type species of the genus Ranavirus (family Iridoviridae). Virology 323:70-84. [DOI] [PubMed] [Google Scholar]
  • 59.Tapiovaara, H., N. J. Olesen, J. Linden, E. Rimaila-Parnanen, and C. H. von Bonsdorff. 1998. Isolation of an iridovirus from pike-perch Stizostedion lucioperca. Dis. Aquat. Organ. 32:185-193. [DOI] [PubMed] [Google Scholar]
  • 60.Tidona, C. A., and G. Darai. 1997. The complete DNA sequence of lymphocystis disease virus. Virology 230:207-216. [DOI] [PubMed] [Google Scholar]
  • 61.Trimble, J. J., S. C. S. Murthy, A. Bakker, R. Grassmann, and R. C. Desrosiers. 1988. A gene for dihydrofolate-reductase in a herpesvirus. Science 239:1145-1147. [DOI] [PubMed] [Google Scholar]
  • 62.Tsai, C. T., J. W. Ting, M. H. Wu, M. F. Wu, I. C. Guo, and C. Y. Chang. 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]
  • 63.Whittington, R. J., C. Kearns, A. D. Hyatt, S. Hengstberger, and T. Rutzou. 1996. Spread of epizootic haematopoietic necrosis virus (EHNV) in redfin perch (Perca fluviatilis) in southern Australia. Aust. Vet. J. 73:112-114. [DOI] [PubMed] [Google Scholar]
  • 64.Williams, T., V. Barbosa-Solomieu, and V. G. Chinchar. 2005. A decade of advances in iridovirus research. Adv. Virus Res. 65:173-248. [DOI] [PubMed] [Google Scholar]
  • 65.Yu, Y. X., M. Bearzotti, P. Vende, W. Ahne, and M. Bremont. 1999. Partial mapping and sequencing of a fish iridovirus genome reveals genes homologous to the frog virus 3 p31, p40 and human eIF2 alpha. Virus Res. 63:53-63. [DOI] [PubMed] [Google Scholar]
  • 66.Zhang, Q. Y., F. Xiao, Z. Q. Li, J. F. Gui, J. H. Mao, and V. G. Chinchar. 2001. Characterization of an iridovirus from the cultured pig frog Rana grylio with lethal syndrome. Dis. Aquat. Organ. 48:27-36. [DOI] [PubMed] [Google Scholar]
  • 67.Zhang, Q. Y., F. Xiao, J. Xie, Z. Q. Li, and J. F. Gui. 2004. Complete genome sequence of lymphocystis disease virus isolated from China. J. Virol. 78:6982-6994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zupanovic, Z., C. Musso, G. Lopez, C. L. Louriero, A. D. Hyatt, S. Hengstberger, and A. J. Robinson. 1998. Isolation and characterization of iridoviruses from the giant toad Bufo marinus in Venezuela. Dis. Aquat. Organ. 33:1-9. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[Supplemental material]

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

RESOURCES