Skip to main content
Journal of Virology logoLink to Journal of Virology
. 2006 Sep;80(17):8439–8449. doi: 10.1128/JVI.00464-06

Genome of Invertebrate Iridescent Virus Type 3 (Mosquito Iridescent Virus)

Gustavo Delhon 1,3,9,*, Edan R Tulman 1,4,5, Claudio L Afonso 1,6, Zhiqiang Lu 1, James J Becnel 2, Bettina A Moser 2,7,8, Gerald F Kutish 1,5,6, Daniel L Rock 1,9
PMCID: PMC1563875  PMID: 16912294

Abstract

Iridoviruses (IVs) are classified into five genera: Iridovirus and Chloriridovirus, whose members infect invertebrates, and Ranavirus, Lymphocystivirus, and Megalocytivirus, whose members infect vertebrates. Until now, Chloriridovirus was the only IV genus for which a representative and complete genomic sequence was not available. Here, we report the genome sequence and comparative analysis of a field isolate of Invertebrate iridescent virus type 3 (IIV-3), also known as mosquito iridescent virus, currently the sole member of the genus Chloriridovirus. Approximately 20% of the 190-kbp IIV-3 genome was repetitive DNA, with DNA repeats localized in 15 apparently noncoding regions. Of the 126 predicted IIV-3 genes, 27 had homologues in all currently sequenced IVs, suggesting a genetic core for the family Iridoviridae. Fifty-two IIV-3 genes, including those encoding DNA topoisomerase II, NAD-dependent DNA ligase, SF1 helicase, IAP, and BRO protein, are present in IIV-6 (Chilo iridescent virus, prototype species of the genus Iridovirus) but not in vertebrate IVs, likely reflecting distinct evolutionary histories for vertebrate and invertebrate IVs and potentially indicative of genes that function in aspects of virus-invertebrate host interactions. Thirty-three IIV-3 genes lack homologues in other IVs. Most of these encode proteins of unknown function but also encode IIV3-053L, a protein with similarity to DNA-dependent RNA polymerase subunit 7; IIV3-044L, a putative serine/threonine protein kinase; and IIV3-080R, a protein with similarity to poxvirus MutT-like proteins. The absence of genes present in other IVs, including IIV-6; the lack of obvious colinearity with any sequenced IV; the low levels of amino acid identity of predicted proteins to IV homologues; and phylogenetic analyses of conserved proteins indicate that IIV-3 is distantly related to other IV genera.


Iridoviruses (IVs) (family Iridoviridae) are large, icosahedral, double-stranded DNA (dsDNA) viruses that cause subclinical to lethal infections in invertebrates and poikilothermic vertebrates. IV particles are 120 to 300 nm in diameter and consist of a central core of nucleic acid and proteins, an intermediate lipid membrane, a T = 147 capsid comprised of copies of the major capsid protein (MCP), and, in virions that bud from the plasma membrane, an external envelope (90, 95). IV genomic DNA is a linear molecule which is circularly permuted and terminally redundant (30). Although IV replication includes nuclear and cytoplasmic stages, virion assembly occurs exclusively in the cytoplasm (28). In contrast to most dsDNA insect viruses, insect IVs are not occluded in a protein matrix.

Based on particle size, host range, DNA cross-hybridization, the presence of a methyl transferase, and the MCP gene sequence, IVs are classified into five genera, with viruses of two genera (Iridovirus and Chloriridovirus) infecting invertebrates, mostly insects (invertebrate iridescent viruses [IIVs]), and viruses of the other three (Ranavirus, Lymphocystivirus, and Megalocytivirus) infecting cold-blooded vertebrates (vertebrate iridoviruses [VIVs]) (47, 84, 91). Complete genome sequences have been determined for viruses representing four of the five genera, including lymphocystis disease virus 1 (LCDV-1) and LCDV-China (LCDV-C) of the genus Lymphocystivirus; tiger frog virus (TFV), frog virus 3 (FV-3), Ambystoma tigrinum virus (ATV), and Singapore grouper iridovirus (SGIV) of the genus Ranavirus; rock bream iridovirus (RBIV) and infectious spleen and kidney necrosis virus (ISKNV) of the genus Megalocytivirus; and IIV-6 (or Chilo iridescent virus) of the genus Iridovirus (22, 34, 35, 42, 43, 69, 73, 76, 98). Also available are the genome sequences of grouper iridovirus (GIV) and orange-spotted GIV (OSGIV), viruses currently lacking taxonomic classification (53, 78). IV genomes range in size from 105 to 212 kbp and contain 96 to 234 largely nonoverlapping open reading frames (ORFs), a G+C content ranging from 27 to 55%, and complex repeat sequences mostly located between coding regions. Genomes exhibit little to no colinearity among genera (see Table 1).

TABLE 1.

Characteristics of IV genomes

Virus Genus Genome size (bp) GC content (%) No. of putative genes Coding density (%) Protein size range (aa) Accession no. Reference
IIV-3 Chloriridovirus 190,132 48 126 68 60-1,377 DQ643392 This work
IIV-6 Iridovirus 212,482 29 234 85 40-2,432 AF303741 42
ATV Ranavirus 106,332 54 96 79 32-1,294 AY150217 43
TFV Ranavirus 105,057 55 105 94 40-1,294 AF389451 35
FV3 Ranavirus 105,903 55 98 80 50-1,293 AY548484 73
SGIV Ranavirus 140,131 48 162 98 41-1,268 AY521625 69
LCDV-1 Lymphocystivirus 102,653 29 110 82 40-1,199 L63545 76
LCDV-C Lymphocystivirus 186,250 27 176 67 40-1,193 AY380826 98
ISNKV Megalocytivirus 111,362 55 124 93 40-1,208 AF371960 34
RBIV Megalocytivirus 112,080 53 118 86 50-1,253 AY532606 22
GIV Unassigned 139,793 49 120 83 62-1,268 AY666015 78
OSGIV Unassigned 112,636 54 121 91 40-1,168 AY894343 53

The type species and currently sole member of the genus Chloriridovirus is Invertebrate iridescent virus type 3 (IIV-3), also known as mosquito iridescent virus (MIV). IIV-3 is characterized by its restricted host range (mosquitoes [Diptera]) and relatively large particle size (180 nm) (17, 90). In contrast, IIVs of the Iridovirus genus have been isolated from diverse hosts of the orders Diptera, Lepidoptera, Hemiptera, and Coleoptera and are approximately 120 nm in diameter (89). Iridovirus species include IIV-1, the first reported IV; IIV-6, the type species of the genus; and several tentative species which remain to be fully characterized (25, 89, 94). Notably, IIV-6 can also infect mosquitoes but causes only sublethal (covert) infections and reduced fitness in Aedes aegypti relative to noninfected conspecifics (55).

IIV-3 was originally isolated from larvae of the salt marsh mosquito Ochlerotatus (formerly Aedes) taeniorhynchus (Wiedemann) (19), with successful transmission to an additional mosquito host, Ochlerotatus sollicitans (93). IIV-3 has been isolated from several other mosquito species, including Aedes vexans, Psorophora ferox, Culiseta annulata, and Culex territans, which are important pests of both humans and domestic animals (14, 17, 93). Early mosquito larval stages are most susceptible to IIV-3 infection, but clinical disease (yellow-green iridescence beneath the epidermis) and death rates are highest in the fourth instar (93). IIV-3 infection of O. taeniorhynchus results in virus replication in the fat body and to a lesser extent in the dermis, imaginal disks, trachea, gonads, and hemocytes (33). Oral and transovarian transmission of IIV-3 have been documented for larval mosquitoes (19, 51, 52, 93). Two IIV-3 strains have been described, a field isolate referred to as regular strain (RMIV) and a laboratory isolate referred to as turquoise strain (TMIV) (93); these cause orange and blue-green iridescence, respectively, in infected larvae.

Despite the role of mosquitoes as significant vectors of human disease, chloriridoviruses are the least studied members of the family Iridoviridae. To date, IIV-3 genomic data are limited to the sequence of the DNA polymerase gene of RMIV (71, 83) and estimations that the TMIV genome contains about twice as much repetitive DNA as that of RMIV (82). Here, we report the complete genomic sequence, with analysis, of IIV-3, RMIV strain. These data, combined with previous work on viruses belonging or related to the other four IV genera, provide a first comparative overview of IV genomics.

MATERIALS AND METHODS

IIV-3 DNA isolation, cloning, and sequencing.

The RMIV strain was originally obtained from field-collected Ochlerotatus taeniorhynchus (Diptera: Culicidae) larvae and maintained by horizontal transmission in the laboratory. RMIV-infected O. taeniorhynchus larvae were ground with a Tekmar Tissuemizer in deionized water and filtered through 400 mesh to remove large insect parts. The filtrate was layered on a continuous HS-40 Ludox gradient and spun for 30 min at 16,000 × g. The band containing RMIV was collected, resuspended in deionized water, and spun for 30 min at 16,000 × g, and genomic DNA was extracted from the pellet as previously described (87). Genomic DNA was incompletely digested with TacI endonuclease (New England Biolabs, Beverly, Mass.), and fragments larger than 1.0 kbp were cloned into the dephosphorylated EcoRI site of pUC19 plasmid vector and grown in Escherichia coli DH10B cells (Gibco-BRL, Gaithersburg, Md.). Plasmids were purified by alkaline lysis according to the manufacturer's instructions (Eppendorf-5 Prime, Boulder, Colo.). DNA templates were sequenced from both ends with M13 forward and reverse primers using dideoxy chain terminator sequencing chemistries (65) and the Applied Biosystems PRISM 3700 automated DNA sequencer (Applied Biosystems, Foster City, Calif.). Chromatogram traces were basecalled with Phred (23) and linearly assembled with Phrap (http://www.phrap.org), with the quality files and default settings used to produce a consensus sequence which was manually edited with Consed (31). An identical sequence was assembled by using the Cap3 assembler with quality files and clone length constraints (38). While the IIV-3 genome assembled essentially in a circular fashion consistent with circular permutation, specifically, randomly located but overlapping termini were observed in independent linear assemblies. For clarity, the IIV-3 genome is presented here in a linear fashion, with overlapping sequence removed from the right terminus of one assembly and the first base of the left terminus arbitrarily numbered base one. The final DNA consensus sequence represented on average 10-fold redundancy at each base position (5,483 reactions) and had a Consed estimated error rate of 0.04/10 kbp.

Sequence analysis.

DNA composition, structure, repeats, and restriction enzyme patterns were analyzed as previously described (1). ORFs longer than 30 codons with a methionine start codon were evaluated for coding potential by the Hexamer (ftp.sanger.ac.uk/pub/rd), Glimmer (64), and Framefinder (http://www.ebi.ac.uk/∼guy) computer programs. Minor ORFs contained within larger ORFs were excluded. Here, 126 ORFs are annotated as potential genes and numbered from left to right. Given the predicted nature of all IIV-3 genes and gene products, ORF names are used throughout this work to indicate both the predicted gene and its putative protein product. DNA and protein comparisons with entries in genetic databases (PROSITE, Pfam, Prodom, Sbase, Blocks, and GenBank) were performed with the BLAST (4), PsiBlast (5), FASTA, TFASTA (63), and HMMER (70) programs. The GCG (20), MEMSAT (44), Psort (61), and SAPS (13) programs were used for general analysis, transmembrane prediction, and physical characterization of predicted proteins. DNA repeats were analyzed with Lalign (http://workbench.sdsc.edu) and Dot Plot (GCG package). Sequence quality, sequence depth, and the lack of obvious polymorphism observed in IIV-3 repeat regions were consistent with that of the rest of the IIV-3 genome assembly. Multiple alignments were performed with Clustal, Dialign-T, and Kalign (49, 72, 75). Phylogenetic analysis was performed with the PHYLO_WIN, TREE-PUZZLE, IQPNNI, PHYLIP, PHYML, and MRBAYES software packages (16, 24, 26, 32, 39, 66), and evolutionary models were selected with ModelGenerator (http://bioinf.nuim.ie/software/modelgenerator). Additional analyses were conducted on alignments where poorly aligned regions were removed with Gblocks (16).

Nucleotide sequence accession number.

The IIV-3 genome sequence has been deposited in the GenBank database under accession no. DQ643392.

RESULTS AND DISCUSSION

Genome organization.

IIV-3 DNA was assembled in a contiguous sequence of 190,132 bp, a figure that differs significantly from previous size estimates based on restriction endonuclease analysis (135,000 bp) and sucrose gradient centrifugation (383,000 bp) (82, 88). IIV-3 nucleotide composition averaged 48% G+C, which is lower than previously reported (54%) (82) and closer to the ranavirus and megalocytivirus genomes than to IIV-6, the other fully sequenced IIV (Table 1). The IIV-3 base composition was not uniformly distributed throughout the genome, with lower G+C values (42%) found in regions containing repeated sequences.

Based on coding potential analysis and similarity to known proteins, 126 IIV-3 ORFs were annotated here (Fig. 1; Table 2). These ORFs, which encode proteins of 60 to 1,377 amino acids (aa), represent a coding density of 68%, one of the lowest among fully sequenced IVs (Table 1). IIV-3 contains 33 unique genes, 27 homologues of genes present in all sequenced IVs, and 52 genes present in IIV-6 but not in VIVs. Consistent with the lack of gene colinearity observed between IIV-6 and VIV and even between LCDV-1 and LCDV-C, viruses likely belonging to the same genus, Lymphocystivirus, the IIV-3 genome exhibited no obvious colinearity with any other completely sequenced IV genome.

FIG. 1.

FIG. 1.

Linear map of the IIV-3 genome. ORFs are numbered from left to right. ORFs transcribed to the right or to the left are located above and below the horizontal line, respectively. Red boxes represent genes encoded by all currently sequenced IV genomes. Blue, yellow, and white boxes are genes unique for IIVs, IIV-3, and selected IVs, respectively. Repetitive DNA sequences are shown as numbered black boxes.

TABLE 2.

IIV-3 ORFs

ORF Nucleotide position Length (codons) Best matcha
Predicted structure and/or functionb
Protein Accession no. Blast score % aa identity
IIV3-001R 2620-3063 148 IIV-6 395R AF303741 97 31
IIV3-002R 3603-4976 458 Drosophila CG4416 product AE014298 122 31
IIV3-003L 5638-5171 156
IIV3-004R 5671-7023 451 IIV-6 067R AF303741 608 39
IIV3-005L 7867-7217 217 Myxococcus unknown AF448145 120 29 Cys-rich protein
IIV3-006R 8063-9544 494 IIV-6 118L AF303741 546 30 Myristylated membrane protein
IIV3-007R 9683-11023 447 TFV 025R AF389451 151 26 FV-3 31-kDa-like early protein
IIV3-008L 12292-11252 347 IIV-6 443R AF303741 172 33
IIV3-009R 12359-15778 1,140 IIV-6 428L AF303741 1977 41 Rpb2
IIV3-010L 17538-15862 559 IIV-6 380R AF303741 582 34 S/T protein kinase
IIV3-011L 19897-17645 751 IIV-6 380R AF303741 287 26
IIV3-012R 20020-21138 373 IIV-6 302L AF303741 171 28 C2H2 Zn finger protein
IIV3-013L 21570-21301 90 IIV-6 141R AF303741 32
IIV3-014L 23106-21709 466 IIV-16 MCP AF025775 1792 58 Major capsid protein
IIV3-015R 23196-23621 142 IIV-22 hypothetical 15.9 kDa P25097 266 53
IIV3-016R 26289-29705 1,139 IIV-6 295L L63545 307 25
IIV3-017R 29876-30742 289 IIV-6 335L AF303741 133 28
IIV3-018L 31615-31121 165 IIV-6 415R AF303741 275 44
IIV3-019R 31881-33098 406 IIV-6 201R AF303741 656 40 Bro family protein
IIV3-020R 33982-34476 165 IIV-6 196R AF303741 255 36 Thioredoxin
IIV3-021L 35147-34530 206 Spodoptera IAP AF186378 139 26 C3HC4 RING finger protein
IIV3-022L 35868-35194 225 TM
IIV3-023R 37466-37783 106
IIV3-024R 38120-39592 491 IIV-6 224L AF303741 828 40 Cysteine protease
IIV3-025R 39626-40159 178 IIV-6 111R AF303741 86 26 TM
IIV3-026R 40206-40883 226 IIV-6 350L AF303741 256 36 VV A1L-like transcription factor
IIV3-027R 40990-41502 171 IIV-6 157L AF303741 111 26 RING finger protein
IIV3-028R 44228-45238 337
IIV3-029R 45281-45859 193 IIV-6 143R AL390114 404 40 Thymidine kinase
IIV3-030L 46385-45954 144 Arabidopsis thaliana AY087324 77 26 Thioredoxin-like protein
IIV3-031R 46406-46831 142 IIV-6 115R AF303741 97 26
IIV3-032R 46927-47763 279 Xenopus unknown BC041236 152 31
IIV3-033L 48462-47881 194 IIV-6 307L AF303741 401 49
IIV3-034R 48593-49390 266 IIV-6 077L AF303741 120 32 Zn finger domain
IIV3-035R 49517-52810 1,098 IIV-6 179R AF303741 1359 34 Protein kinase active-site signature
IIV3-036R 56099-56872 258 IIV-6 219L AF303741 245 34
IIV3-037L 57696-56938 253 IIV-6 060L AF303741 148 27 TM
IIV3-038R 57775-59412 546 IIV-6 098R AF303741 515 28
IIV3-039R 59602-60930 443 IIV-6 393L AF303741 314 28 FV-3 IE ICP46-like
IIV3-040R 61046-61294 83
IIV3-041R 61328-61714 129 IIV-6 453L AF303741 172 35 Thioredoxin
IIV3-042R 61812-62288 159 IIV-6 136R AF303741 280 45 Zn binding signature
IIV3-043R 62326-62517 64 IIV-6 010R AF303741 169 53
IIV3-044L 63715-62711 335 Saccharomyces YOLO44W Z74786 212 29 S/T protein kinase
IIV3-045R 63932-64201 90
IIV3-046R 64376-66172 599 IIV-6 229L AF303741 418 27
IIV3-047R 66434-67654 407 IIV-6 337L AF303741 583 37 Virion-associated membrane protein
IIV3-048L 68853-67726 376 IIV-6 376L AF303741 763 46 Ribonucleotide reductase small subunit
IIV3-049R 69009-71057 683 CPXV 158 AF482758 175 24
IIV3-050L 71580-71125 152 IIV-6 145L AF303741 110 25
IIV3-051L 75421-74576 282 IIV-6 213R AF303741 109 35
IIV3-052L 77801-75534 756 IIV-6 205R AF303741 559 31 NAD-dependent DNA ligase
IIV3-053L 78645-78220 142 Encephalitozoon Rbp 19 kD AL590447 73 25 Rpb7
IIV3-054L 83584-82658 309
IIV3-055R 83754-84164 137 IIV-6 349L AF303741 184 33 Transcription factor SII
IIV3-056L 85251-84220 344 IIV-6 287R AF303741 319 27
IIV3-057L 85749-85360 130
IIV3-058R 85937-86386 150 IIV-6 391R AF303741 233 32
IIV3-059L 88441-86762 560 IIV-6 012L AF083915 582 34 Exonuclease II
IIV3-060L 89678-88827 284 PBCV-1 A193L U42580 130 25 PCNA-like protein
IIV3-061R 94254-95717 488 IIV-6 467R AF303741 210 25
IIV3-062L 96241-95546 232 ATV hypothetical protein AY150217 152 56 Repetitive protein
IIV3-063R 96289-96924 212 IIV-6 309L AF303741 80 32
IIV3-064L 97827-96988 280
IIV3-065R 97890-99779 630 Rickettsia RR alpha chain AE008624 1167 57 Ribonucleotide reductase large subunit
Protein Accession no. Blast score % aa identity
IIV3-066L 100438-99860 193 IIV-6 357R AF303741 97 25 TM
IIV3-067L 102380-101661 240 IIV-6 197R AF303741 262 44 S/Y protein phosphatase
IIV3-068R 102821-103423 201 IIV-6 401R AF303741 404 44 HMG-like
IIV3-069L 104770-103502 423 IIV-6 198R AF303741 407 30
IIV3-070L 105749-104904 282 IIV-6 306R AF303741 171 31 SWIB/MDM2 domain
IIV3-071L 106597-105935 221 IIV-6 259R AF303741 165 27
IIV3-072L 107142-106675 156 IIV-6 374R AF303741 103 28
IIV3-073R 107238-107768 177 IIV-6 234R AF303741 93 26 TM
IIV3-074L 110351-107850 834 IIV-6 268L AF303741 496 30
IIV3-075R 112290-112763 158 Debaryomyces plasmid protein AJ011124 132 35 DNA Pol III box
IIV3-076L 114217-113078 380 IIV-6 369L AF303741 466 36 Xeroderma pigmentosum protein
IIV3-077R 115523-115963 147
IIV3-078R 116034-117074 347 IIV-6 244L AF303741 691 49 Phosphoesterase
IIV3-079L 118306-117122 395 IIV-6 282R AF303741 688 43
IIV3-080R 118504-119226 241 CMLV112R AF438165 136 28 MutT-like protein
IIV3-081L 120310-119786 175
IIV3-082L 120800-120315 162
IIV3-083L 121653-121078 192 IIV-6 358L AF303741 87 28
IIV3-084L 124239-121708 844 IIV1 98-kDa protein P22856 1096 38 OTU domain
IIV3-085L 125066-124584 161 IIV-6 203L AF303741 106 22 TM
IIV3-086L 128555-125217 1,113 IIV-6 045L AF083915 2177 42 DNA topoisomerase II
IIV3-087L 134049-131110 980 IIV-6 022L AF083915 1127 31 NTPase SNF2
IIV3-088R 134089-134871 261 IIV-6 075L AF303741 630 47 ATPase
IIV3-089L 135221-134928 98
IIV3-090L 139445-135315 1,377 IIV-6 176R AF303741 1723 41 Rpb1
IIV3-091L 142900-139613 1,096 IIV-6 443R AF303741 945 34 Repetitive protein
IIV3-092R 142975-143487 171 IIV-6 454R AF303741 34 Rpb5
IIV3-093L 144466-143564 301 IIV-6 019R AF303741 271 28
IIV3-094L 147069-144583 829 IIV-6 050L AF303741 382 27 P-loop nucleoside triphosphatase
IIV3-095L 148265-147177 363 Mouse type 2 metalloprotease D86332 375 40 Metalloprotease
IIV3-096R 148391-148834 148 IIV-6 347L AF303741 221 42 Erv1/Air-like thiol oxidoreductase
IIV3-097L 150653-150039 205 IIV-6 170L AF303741 191 30 RuvC-like resolvase
IIV3-098L 152364-150805 520 IIV-6 439L AF303741 535 33 Protein kinase-like protein
IIV3-099R 155290-156675 462 IIV-6 329R AF303741 314 41
IIV3-100L 157359-156724 212 IIV-6 378R AF303741 460 44
IIV3-101R 157391-158389 333 IIV-6 142R AF303741 431 39 Ribonuclease III
IIV3-102R 161028-161390 121 Buchnera pyrophosphorylase AE014078 95 38
IIV3-103L 161631-161452 60
IIV3-104L 162248-161691 186 IIV-6 355R AF303741 390 47 CTD phosphatase
IIV3-105R 162408-163145 246 IIV-6 359L AF303741 90 29
IIV3-106R 163297-164715 473 IIV-6 030L AF083915 817 42 Helicase
IIV3-107R 164859-165644 262 IIV-6 117L AF303741 222 33 TM
IIV3-108L 166065-165754 104 IIV-6 161L AF303741 311 57 Helicase, C-terminal domain
IIV3-109L 167108-166122 329 IIV-6 161L AF303741 444 39 Helicase, N-terminal domain
IIV3-110R 167222-167992 257
IIV3-111R 168970-169476 169 IIV-6 414L AF303741 184 34 MutT-like protein
IIV3-112R 169526-169861 112 IIV-6 466R AF303741 95 33 TM
IIV3-113L 172363-169943 807 IIV-6 155L AF303741 300 32
IIV3-114L 172916-172482 145 Cyprinus carpio MMp AB057407 91 40 Metalloprotease prodomain
IIV3-115R 173073-173300 76 IIV-6 342R AF303741 142 50
IIV3-116R 173407-174168 254
IIV3-117L 174511-174308 68
IIV3-118R 174533-174754 74
IIV3-119R 174807-175184 126 Pneumocystis endoprotease AF009223 140 42 Proline-rich protein
IIV3-120R 179618-183040 1,141 IIV-3 DNA polymerase AJ312708 4702 89 DNA polymerase
IIV3-121R 183279-186101 941 IIV-6 184R AF303741 1367 34 VV D5R NTPase/helicase
IIV3-122R 186194-186394 67
IIV3-123L 186865-186449 139
IIV3-124R 186910-187617 236 Xanthomonas ribonuclease AE011846 106 27
IIV3-125R 187711-188601 297
IIV3-126R 188648-188962 105
a

CPXV, cowpox virus; CMLV, camelpox virus. Accession numbers are numbers of homologous sequences from the GenBank database.

b

Function was deduced from the degree of similarity to known genes and from Prosite signatures.TM, transmembrane.

Repeated sequences.

The IIV-3 genome contained highly repetitive DNA sequences in 15 distinct regions (designated R1 to R15 from left to right) that were 0.8 to 4.6 kbp in length, irregularly distributed, and apparently noncoding (Fig. 1). IIV-3 DNA repeats were complex, mostly direct and imperfect, and they could be divided into two groups based on sequence similarity, with group I repeats located at R2, R4, R6 to R8, R11 to R13, and R15 and group II repeats located at R1, R3, R5, R9, R10, and R14. Group I repeats were comprised of a 100-nucleotide (nt) sequence present in 2 to 10 copies which had 83 to 100% nucleotide identity. Group II repeats were also comprised of a sequence of approximately 100 nt and were present in two to six copies which had 80 to 99% nucleotide identity. Group I and group II repeats had only 62 to 68% nucleotide identity. Repeated sequences in each group contained invariant motifs (e.g., TAAATTTC, AATC, and GCAT in group I repeats; GAGTT, ATGCGT, and GAAATTT in group II repeats) flanked by less-conserved sequences.

DNA repeats are present in the intergenic regions of all fully sequenced IVs; however, their extent, arrangement, localization, and repeated sequence motifs differ between genera (22, 34, 35, 42, 43, 53, 69, 73, 76, 78, 98). In contrast to previous reports, IIV-3 repeated sequences were extensive (20% of the genome), resembling IIV-9 and IIV-16, in which 25% and 39% of the genomes, respectively, are made up of repetitive DNA (10, 58, 89). Although late transcription has been detected from IIV-9 repeated DNA regions, the role of IV DNA repeats is unknown (59). Repeated sequences are known to be involved in genome replication and gene transcription in other DNA viruses. For example, in nucleopolyhedroviruses (NPVs), repetitive DNA sequences function as enhancers of transcription and as origins of genomic DNA replication (3, 74).

Notable IIV-3 genes. (i) Viral transcription and DNA replication.

The IIV-3 genome contained several genes with predicted roles in viral transcription and RNA processing, including five RNA polymerase II subunit (Rpb) homologues, and in viral DNA replication, metabolism, and maintenance (Table 2).

IIV3-053L is a novel IV protein with limited similarity to Rpb7, a DNA-dependent RNA polymerase II subunit highly conserved among eukaryotes and having roles in DNA repair, transcription, and RNAi-directed chromatin silencing (18, 21, 56) (Fig. 2). Cellular Rpb7 forms a complex with Rpb4 near the transcript exit groove and the C-terminal domain (CTD) linker region of RNA polymerase II, interacts with various transcription factors and CTD phosphatase FCP1, and is essential for cell viability in yeast (18). IIV3-053L was most similar to Encephalitozoon cuniculi Rpb7 (25% amino acid identity over 142 amino acids) (Fig. 2), particularly in the N-terminal half of the protein and including G64 in the predicted tip loop involved in binding of Rpb7 to the RNA polymerase II core (6). Less conserved was the C-terminal half of IIV3-053L, which lacked 20 aa and a predicted three-stranded antiparallel β-sheet relative to E. cuniculi Rpb7 (6, 77). IIV3-053L lacked recognizable Rpb7 Pfam signatures, suggesting that it, like homologues in Giardia, Methanosarcina, and mimivirus, is a highly divergent Rpb7 whose function in transcription remains to be determined.

FIG. 2.

FIG. 2.

Multiple amino acid sequence alignment of IIV3-053L with selected Rpb7 proteins. Asterisks, colons, and single dots denote identical, conserved, and semiconserved residues, respectively (74). Rpb7 protein sequence names correspond to the following accession numbers: mosquito (Anopheles gambiae), XM_320916; rat, NM_053948; Saccharomyces cerevisiae, NC_001136; E. cuniculi, NC_003234; mimivirus (Acanthamoeba polyphaga mimivirus), NC_006450.

IIV3-092R had limited similarity with predicted Rpb5-like proteins encoded by African swine fever virus (ASFV), mimivirus, and several IVs. Differences between IIV3-092R and other viral Rpb5-like proteins were greatest in the N-terminal two-thirds of the protein, a region in cellular Rpb5 proposed to interact with transcriptional regulators (85).

IIV3-104L was similar to the FCP1 phosphatase catalytic domain (FCPH), with orthologues in all fully sequenced IVs. FCP1 is responsible for dephosphorylating Rpb1 CTD and, together with specific kinases, driving the transition of RNA polymerase II from the initiation to elongation modes. Like SCP1, a recently described FCP1-like phosphatase with similar roles in transcription regulation, IIV3-104L and IV homologues lack most sequences N-terminal of the FCPH domain, including BRCT and TFIIF binding domains (96). Given the lack of recognizable CTD heptapeptide repeats in IV Rpb1 homologues, IV IIV3-104L-like proteins may conceivably target host Rpb1, a protein which is involved in the nuclear phase of IV transcription (29). Whether IIV3-104L-like proteins represent the IV virion-associated protein (VATT) proposed to modify host RNA polymerase II and to initiate iridoviral immediate-early transcription remains to be determined (92).

IIV3-080R and IIV3-111R contained domains characteristic of MutT-like proteins, including the signature motif [G-X5-E (D in IIV3-080R)-X4-5-C/T-L/A-X-RE-F/L-X-EE-X-G/T] at positions 136 to 157 and 76 to 98, respectively. The MutT (or nudix) domain is found in certain phosphohydrolases which are believed to eliminate toxic nucleotide derivatives and to regulate the levels of signaling nucleotides in both eukaryotes and prokaryotes (57). While IIV3-111R had additional similarity with IIV-6 and LCDV MutT-like proteins, IIV3-080R is a novel IV ORF more similar to poxvirus MutT proteins (28% amino acid identity to variola minor virus F10R over 169 aa) (Table 2). A poxvirus MutT-like protein, vaccinia virus (VV) D10R, is essential for virus infectivity and appears to function as a repressor of host and viral transcription and translation (67). Whether MutT proteins play a similar role in IV remains to be determined.

IIV3-060L had similarity with viral homologues of proliferating cell nuclear antigen (PCNA). While most similar to the Paramecium bursaria chlorella virus (PBCV-1) PCNA-like protein A193L (25% amino acid identity, over 239 aa), IIV3-060L was less similar to PCNA-like proteins encoded by all currently sequenced IVs (17 to 21% amino acid identity) and by NPVs and mimivirus. Cellular PCNA, an acidic protein (pI 4.5) with critical roles in DNA replication and repair, assembles as a ring-shaped homotrimer around dsDNA and is highly conserved among eukaryotes. Nine R and K residues within each monomer confer a net positive electrostatic potential to the central channel, a feature thought to be important for PCNA sliding clamp function (48). In contrast, IV PCNA-like proteins are highly divergent between genera, have predicted pI values ranging from 6.4 (IIV-6) to 9.2 (IIV-3) and, with the exception of conserved residues K37, R102, K168, and K231 in IIV3-060L, contain a pattern and distribution of positive charges distinct from that of cellular PCNA. VV G8R, one of the most divergent viral PCNA-like proteins, is an essential protein that affects late transcription rather than viral DNA replication (46). While the function of IV PCNA-like proteins is unknown, they also may have a novel role during viral replication.

(ii) IIV-3 genes lacking VIV homologues.

IIV-3 shared with IIV-6 52 genes absent in VIVs (Fig. 1). These potentially IIV-specific genes encoded several proteins with predicted functions involving DNA replication and maintenance, including DNA topoisomerase IIA (IIV3-086L), Pif1-like SFI helicase (IIV3-106R), NAD-dependent DNA ligase (IIV3-052L), HMGB-like protein (IIV3-068R), and Swib/mdm2 homology domain-containing protein (IIV3-070L), and several predicted to function in protein modification, including OTU-like cysteine protease (IIV3-084L), type 2 metalloprotease (IIV3-095L), and RING finger proteins (IIV3-021L and IIV3-027R). Additionally, a protein containing the BRO motif found in proteins encoded by a range of other insect viruses (IIV3-019R) was identified.

IIV-specific proteins predicted to manipulate DNA contained novel features relative to cellular or viral homologues. IIV3-086L, similar to topoisomerase IIA encoded by other viruses, lacked the C-terminal domain important for nuclear targeting and for interaction with other proteins (50). Similarly, IIV3-106R contained all seven motifs (I, Ia, and II to VI) characteristic of cellular Pif1-like helicases but lacked most of the N-terminal sequences implicated in Pif1/chromatin factor interaction (reviewed in reference 8). The IIV3-052L NAD-dependent DNA ligase contained a putative C-terminal BRCT domain (position 648 to 720) absent in entomopoxviral homologues, suggesting that IIV DNA ligases, similar to cellular homologues, might direct multimeric complex assemblies through this domain (27).

IIV3-068R contained A and B boxes (positions 70 to 138 and 143 to 185, respectively) similar to those found in non-histone chromatin proteins of the high mobility group B (HMGB), proteins that bind and distort DNA and interact with a number of transcription factors and DNA repair and recombination proteins (reviewed in reference 2). IIV HMGB-like proteins lack the acidic tail found in cellular HMGBs. Interestingly, a cellular HMGB mutant lacking the acidic tail exhibits 100-fold-higher DNA binding affinity than the wild-type protein, leading to a block in nucleosome sliding (11). Iridovirus HMGB may conceivable play a structural role in IIV genome conformation.

Several IIV-specific proteins shared similarity or functional motifs with proteins involved in modification of other viral or host proteins. Similar to cellular and viral members of the OTU (for Drosophila “ovarian tumor gene”) cysteine protease superfamily, IIV3-084L contained in its C terminus the four conserved motifs of an OTU homology domain (position 596 to 719), including residues in motif I (C601) and motif IV (H718) thought to be important for cleavage (54, 68). IIV3-084L also contained a highly positively charged region (position 137 to 218) and a putative bipartite NLS (position 410), features shared by IIV-1 (Tipula iridescent virus) late protein L96 (35% amino acid identity) and IIV-6 232R (22% amino acid identity) (37). The classification of OTU proteins as cysteine proteases was originally based solely on similarity to the arterivirus NSP2 cysteine protease; however, a number of deubiquitinating enzymes exhibiting proteolytically active OTU domains, including a potent downregulator of NF-κB, were recently described (7, 12, 54, 86). IIV3-084L may similarly affect ubiquitin-mediated protein degradation to counteract antiviral cellular responses.

IIV3-095L, a protein with homologues in viruses infecting insects, including IIV-6, granuloviruses, and entomopoxviruses, was similar to mammalian type 2 matrix metalloproteinases (MMp). IIV3-095L contained an N-terminal signal peptide, the MMp prodomain consensus sequence (or cysteine switch) (PRCXXPD [position 117 to 123]), the catalytic domain signature (HEXXHXXGXXH [position 275 to 285]), a conserved Met at position 293 (reviewed in reference 62), and RRRR and RTRR motifs similar to the equivalently located RRKR furin cleavage target motif of mammalian MMp, indicating that IIV3-095L is likely activated by furin-mediated proteolysis of the prodomain (81). IIV3-095L lacked predicted transmembrane domains and the hemopexin-like domain which affects substrate recognition in many mammalian MMps. Relative to IIV3-095L, the MMp homologue in IIV-6 (165R) lacks 100 N-terminal amino acids, including the signal peptide, the cysteine switch signature, and the potential furin cleavage site, suggesting differences in compartmentalization and activation mechanisms between IIV MMps. IIV3-095L could conceivably function as is the case for entomopathogenic bacterial proteins, targeting receptors, antimicrobial peptides, or other factors involved in insect responses to pathogens, thus facilitating virus replication. Alternatively, viral MMp could act on the extracellular matrix (e.g., peritrophic matrix of lepidopterous larvae), facilitating virus spread within the host (80). Notably, an MMp-like prodomain is also present in IIV3-114L; however, lack of the prodomain Cys residue critical for MMp function and other MMp features make IIV3-114L an unlikely MMp.

IIV3-021L and IIV3-027R contained C-terminal C3HC4-type RING finger motifs which are characteristic of a diverse array of single-subunit E3 ubiquitin ligases and critical for E3 recruitment of ubiquitin-conjugating enzymes to specific substrates (36). Given the importance of the RING domain in cellular E3 ubiquitin ligase function, its presence in IIV3-021L and IIV3-027R suggests that they too may affect protein ubiquitination to mediate specific virus-host interactions. Notably, the IIV3-021L RING motif (CX2CX4RX5PCXHX3CX2CX9CPXC) exhibits particular similarity to the RING motif present in the E3-like baculovirus inhibitor of apoptosis protein 3 (IAP-3) and is preceded by a single, truncated domain (CXCX10EX5H [position 126 to 145]) similar to the baculovirus IAP repeat (BIR) domain involved in caspase inhibition by viral and mammalian IAPs (60). These features suggest that IIV3-021L may also participate in caspase inhibition to affect apoptosis in the host cell.

IIV3-019R contained an N-terminal baculovirus repeated ORF (BRO)-family homology domain (Bro-N domain, position 24 to 132), 62% identical to the IIV-6 201R Bro-N domain, and included a single-stranded DNA binding motif at position 24 to 57. The C terminus of IIV3-019R, however, is most similar to the C-terminal domains of proteins lacking Bro-N domains (entomopoxvirus AMV207 and AMV209; 27% amino acid identity), consistent with the extensive domain shuffling found in BRO family proteins (40). BRO family proteins are encoded, often in multiple copies, by many invertebrate dsDNA viruses and by bacteriophages. While three BRO genes have been identified in IIV-6, IIV-3, like IIV-31, contains only one (9, 41). Bombyx mori NPV BRO proteins (BRO-a to BRO-e) are expressed early during infection, with BRO-a and BRO-c proteins interacting with chromatin (45, 97). The requirement of BRO proteins for virus replication seems to vary with the specific virus-host system. Disruption of the single BRO gene in the Autographa californica multicapsid NPV did not affect virus replication or virus pathogenicity in instar larvae (9). Conversely, inability to recover B. mori NPV BRO-d deletion mutants suggested a function essential for virus replication (45).

(iii) IIV-3 genes encoding conserved IV domains.

IIV3-016R, IIV3-033L, IIV3-056L, and IIV3-107R are proteins of unknown function similar to proteins encoded by all currently sequenced IVs. IIV3-016R contained in the central region of the protein a conserved 90-aa domain which included the motif YXCX8-9GX3NX11-12PCCY (position 502 to 533), also found in entomopoxvirus proteins (MSV063 and AMV105) which have similarity with VV A7L early transcription factor subunit. IIV3-033L contained a putative N-terminal transmembrane domain and a 100-aa domain also present in proteins of unknown function encoded by ascovirus (accession no. CAC19143) and Symbiobacterium thermofilum, an uncultivable bacterium that depends on bacterial commensalism for growth (79). IIV3-056L and IIV3-107R, predicted to contain an N-terminal transmembrane domain, shared with IV homologues a 50-aa domain which included the motifs F/Y-X4-V/I-R-G-X11-A-X2-h-h-X13-14-G-X-P-X-P (position 144 to 187, where “h” is hydrophobic) and R-X5-D-P/F-IRGD-L/V-X-I-X-P-X5-F-X5-P-X3-L-X2-G (position 116 to 151), respectively. While conserved, the functional relevance of these protein domains is unknown.

Comparison between IIV-3 and other IV genomes.

IIV-3 overall resembled other IVs in genome size, DNA composition, and gene complement. Data also indicated that among completely sequenced IVs, IIV-3 most closely resembled IIV-6, the only other sequenced IIV. Despite these similarities, IIV-3 contained novel genomic features that indicate its distant relationship to other IV genera and confirm its unique position within the family Iridoviridae.

IIV-3 contained homologues of 27 genes present in all currently sequenced IVs (Fig. 1, red boxes), indicating that these genes play critical and likely essential roles in aspects of IV biology. Approximately half of the conserved genes are involved in viral transcription/DNA replication (IIV3-009R, IIV3-029R, IIV3-048L, IIV3-055R, IIV3-060L, IIV3-076L, IIV3-087L, IIV3-090L, IIV3-101R, IIV3-104L, IIV3-120R, and IIV3-121R). Others encode homologues of protein kinase, MCP, FV3 IE ICP46-like protein, ATPase, and Erv1/Alr-like protein (IIV3-010L, IIV3-014L, IIV3-039R, IIV3-088R, and IIV3-096R, respectively). The functions of nine IIV-3 genes conserved among IVs, including two genes encoding putative transmembrane proteins (IIV3-006R and IIV3-107R), are unknown.

Fifty-two IIV-3 genes have IIV-6 but not VIV counterparts (Fig. 1, blue boxes), suggesting that these genes (IIV-3/IIV-6 genes) function in infection of invertebrate hosts. Although approximately a third of the IIV-3/IIV-6 genes were clustered in three regions in the IIV-3 genome (positions 29876 to 35147, 99860 to 106597, and 143564 to 157359), even these regions lacked discernible colinearity. Only 12 of 52 IIV-3/IIV-6 genes have a predicted function/activity, with 7 genes likely involved in DNA replication/maintenance or gene expression (IIV3-026R, IIV3-052L, IIV3-059L, IIV3-068R, IIV3-070L, IIV3-078R, IIV3-086L, and IIV3-111R), 4 genes encoding protein modification enzymes (IIV3-020R, IIV3-067L, IIV3-095L, and IIV3-098L), and 1 gene encoding an apoptosis regulator (IIV3-021L). Of the IIV-3/IIV-6 genes with unknown function, six encode predicted transmembrane proteins (IIV3-025R, IIV3-037L, IIV3-066L, IIV3-073R, IIV3-085L, and IIV3-112R), three encode Zn finger proteins (IIV3-012R, IIV3-027R and IIV3-034R), and one encodes a Bro-like protein (IIV3-019R). An additional similarity between IIV-3 and IIV-6 was the lack of specific genes present in VIVs, including those encoding DNA methyltransferase and proteins proposed to play roles in evading host responses (e.g., β-OH steroid oxidoreductase, eIF-2α, caspase recruitment domain-containing protein). Notably, the only homologue of IIV3-015R is a gene similarly located adjacent to the MCP gene in IIV-22, an IIV infecting the dipteran Simulium variegatum (15).

While certain features of the IIV-3 genome were most similar to IIV-6, others made clear the unique nature of IIV-3 and its distant relationship to IIV-6. These included differences in the nature and extent of repetitive DNA, genome composition, gene colinearity, and gene complement. Thirty-three predicted IIV-3 proteins are not encoded by other IVs (Fig. 1, yellow boxes) and, with the exception of IIV3-044L (protein kinase), IIV3-053L (Rpb7), and IIV3-080R (MutT-like protein), they lacked similarity to any other protein. VIV gene homologues present in either IIV-3 or IIV-6, but not both, include IIV3-007R, which encodes a homologue of the FV-3 IE 31-kDa protein absent in IIV-6, and thymidylate synthase and dUTPase genes, present in IIV-6 and many VIVs but absent in IIV-3. While predicted IIV-3 proteins were most similar to IIV-6 homologues in pairwise comparisons, amino acid identities ranged only from 22 to 53%. Finally, phylogenetic analyses clearly separated IIV-3 from IIV-6 (Fig. 3). Analysis of a large, concatenated protein data set indicated that IIV-3 and IIV-6, while grouping together within the Iridoviridae, have genetic distances comparable to those between other IV genera, consistent with their classification into separate genera (Fig. 3A). Analysis of available IIV MCP data also indicated a clear separation between IIV-3 and IIV-6 and indicated that IIV-3, while unique, was more similar to a group of viruses which included IIV-22, consistent with IIV-3/IIV-22 colinearity at the IIV3-015/016 locus (Fig. 3B). Notably, several other members of this group (IIV-1, IIV-9, IIV-16) are currently classified as members of the genus Iridovirus. Given that here IIV-1, IIV-9, and IIV-16 were as phylogenetically distinct from IIV-6 (genus Iridovirus) as was IIV-3 (genus Chloriridovirus), reconsideration of current IIV taxonomy may be in order (Fig. 3B).

FIG. 3.

FIG. 3.

Phylogenetic analysis of IIV-3 proteins (A) and IIV MCP (B). (A) Eleven conserved IIV-3 proteins (IIV3-009R, IIV3-014L, IIV3-048L, IIV3-055R, IIV3-076L, IIV3-087L, IIV3-088R, IIV3-090L, IIV3-101R, IIV3-120R, and IIV3-121R) were concatenated and aligned with similar data sets from other iridoviruses by using Kalign. The unrooted tree for 9,190 aligned characters was generated by the maximum likelihood tree search strategy, including the WAG model for correction for multiple substitutions, the four-category discrete gamma model for correction for among-site rate variation, 100 bootstrap replicates, and default settings, as implemented in Phyml. Bootstrap values greater than 70 are indicated at the appropriate nodes, and dots indicate values of 100. Sequences from the following viruses and accession numbers were compared: IIV-3, DQ643392; IIV-6, AF303741; LCDV-1, L63545; LCDV-C, AY380826; GIV, AY666015; SGIV, AY521625; FV-3, AY548484; ATV, AY150217; TFV, AF389451; ISKNV, AF371960; RBIV (strain KOR-TY1), AY532606; OSGIV, AY894343. (B) IIV3-014L was aligned with available IIV MCP sequences and LCDV MCP as an outgroup with Clustal, and the tree was generated as described above. Sequences from the following viruses and accession numbers were compared: IIV-24, AF042340; IIV-30, AF042336; IIV-29, AF042339; AgIV, AF042343; IIV-2, AF042335; IIV-9, AF025774; IIV-23, AF042342; IIV-22, M32799; IIV-1, M33542; IIV-16, AF025775; IIV-3, DQ643392; IIV-31, AF042337; PjIV, AF042338; IIV-6, AF303741; LCDV, AY849392. Scales indicate estimated distances. Similar topologies were obtained with maximum likelihood as implemented in MRBAYES and/or TREE-PUZZLE and with neighbor-joining and maximum-parsimony algorithms as implemented in PHYLO_WIN.

Conclusions.

The genome sequence of IIV-3, sole member of the genus Chloriridovirus, has been determined. IIV-3 contains a number of novel genes which are likely important for infection of the natural mosquito host. Comparison between IIV-3, IIV-6, and available non-IIV-6 IIV gene sequences suggests that diversity within IIV may be greater than previously recognized. Genes common to all four IV genera, which include those encoding enzymes involved in RNA transcription, DNA replication, and protein processing, likely define the genetic core of the family Iridoviridae.

Acknowledgments

We thank A. Lakowitz and A. Zsak for providing excellent technical assistance.

REFERENCES

  • 1.Afonso, C. L., E. R. Tulman, Z. Lu, E. Oma, G. F. Kutish, and D. L. Rock. 1999. The genome of Melanoplus sanguinipes entomopoxvirus. J. Virol. 73:533-552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Agresti, A., and M. E. Bianchi. 2003. HMGB proteins and gene expression. Curr. Opin. Genet. Dev. 13:170-178. [DOI] [PubMed] [Google Scholar]
  • 3.Ahrens, C. H., M. N. Pearson, and G. F. Rohrmann. 1995. Identification and characterization of a second putative origin of DNA replication in a baculovirus of Orgyia pseudotsugata. Virology 207:572-576. [DOI] [PubMed] [Google Scholar]
  • 4.Altschul, S. F., and D. J. Lipman. 1990. Protein database searches for multiple alignments. Proc. Natl. Acad. Sci. USA 87:5509-5513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Altschul, S. F., A. A. Schaffer, J. Zhang, Z. Zhang, W. Miller, and D. J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25:3389-3402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Armache, K. J., S. Mitterweger, A. Meinhart, and P. Cramer. 2005. Structures of complete RNA polymerase II and its subcomplex, Rpb4/7. J. Biol. Chem. 280:7131-7134. [DOI] [PubMed] [Google Scholar]
  • 7.Balakirev, M. Y., S. O. Tcherniuk, M. Jaquinod, and J. Chroboczek. 2003. Otubains: a new family of cysteine proteases in the ubiquitin pathway. EMBO Rep. 4:517-522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bessler, J. B., J. Z. Torres, and V. A. Zakian. 2001. The Pif1 subfamily of helicases: region-specific DNA helicases? Trends Cell Biol. 11:60-65. [DOI] [PubMed] [Google Scholar]
  • 9.Bideshi, D. K., S. Renault, K. Stasiak, B. A. Federici, and Y. Bigot. 2003. Phylogenetic analysis and possible function of bro-like genes, a multigene family widespread among large double-stranded DNA viruses of invertebrates and bacteria. J. Gen. Virol. 84:2531-2544. [DOI] [PubMed] [Google Scholar]
  • 10.Bigot, Y., K. Stasiak, F. Rouleux-Bonnin, and B. A. Federici. 2000. Characterization of repetitive DNA regions and methylated DNA in ascovirus genomes. J. Gen. Virol. 81:3073-3082. [DOI] [PubMed] [Google Scholar]
  • 11.Bonaldi, T., G. Langst, R. Strohner, P. B. Becker, and M. E. Bianchi. 2003. The DNA chaperone HMGB1 facilitates ACF/CHRAC-dependent nucleosome sliding. EMBO J. 21:6865-6873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Borodovsky, A., H. Ovaa, N. Kolli, T. Gan-Erdene, K. D. Wilkinson, H. L. Ploegh, and B. M. Kessler. 2002. Chemistry-based functional proteomics reveals novel members of the deubiquitinating enzyme family. Chem. Biol. 9:1149-1159. [DOI] [PubMed] [Google Scholar]
  • 13.Brendel, V., P. Bucher, I. R. Nourbakhsh, B. E. Blaisdell, and S. Karlin. 1992. Methods and algorithms for statistical analysis of protein sequences. Proc. Natl. Acad. Sci. USA 89:2002-2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Buchatsky, L. P. 1977. An iridovirus from larvae of Culisetta annulata and Culex territans. Acta Virol. 21:85-86. [PubMed] [Google Scholar]
  • 15.Cameron, I. R. 1990. Identification and characterization of the gene encoding the major structural protein of insect iridescent virus type 22. Virology 178:35-42. [DOI] [PubMed] [Google Scholar]
  • 16.Castresana, J. 2000. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol. Biol. Evol. 17:540-552. [DOI] [PubMed] [Google Scholar]
  • 17.Chapman, H. C., T. B. Clark, and D. B. Woodard. 1966. Additional mosquito hosts of the mosquito iridescent virus. J. Invertebr. Pathol. 8:545-546. [DOI] [PubMed] [Google Scholar]
  • 18.Choder, M. 2004. Rpb4 and Rpb7: subunits of RNA polymerase II and beyond. Trends Biochem. Sci. 29:674-681. [DOI] [PubMed] [Google Scholar]
  • 19.Clark, T. B., W. R. Kellen, and P. T. M. Lum. 1965. A mosquito iridescent virus (MIV) from Aedes taeniorhynchus (Wiedemann). J. Invertebr. Pathol. 7:519-521. [DOI] [PubMed] [Google Scholar]
  • 20.Devereux, J., P. Haeberly, and O. Smithies. 1984. A comprehensive set of sequence analysis programs for the VAX. Nucleic Acids Res. 12:387-395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Djupedal, I., M. Portoso, H. Spahr, C. Bonilla, C. M. Gustafsson, R. C. Allshire, and K. Ekwall. 2005. RNA Pol II subunit Rpb7 promotes centromeric transcription and RNAi-directed chromatin silencing. Genes Dev. 19:2301-2306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.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]
  • 23.Ewing, B., L. Hillier, M. C. Wendl, and P. Green. 1998. Base-calling of automated sequencer traces using phred. I. Accuracy assessment. Genome Res. 8:175-185. [DOI] [PubMed] [Google Scholar]
  • 24.Felsenstein, J. 1989. PHYLIP—phylogeny inference package (version 3.2). Cladistics 5:164-166. [Google Scholar]
  • 25.Fukaya, M., and S. Nasu. 1966. A Chilo iridescent virus from the rice stem borer, Chilo suppresalis Walker (Lepidoptera:Pyralidae). Appl. Entomol. Zool. 1:69. [Google Scholar]
  • 26.Galtier, N., M. Gouy, and C. Gautier. 1996. SEAVIEW and PHYLO WIN: two graphic tools for sequence alignment and molecular phylogeny. Comput. Appl. Biol. Sci. 12:543-554. [DOI] [PubMed] [Google Scholar]
  • 27.Glover, J. N., R. S. Williams, and M. S. Lee. 2004. Interactions between BRCT repeats and phosphoproteins: tangled up in two. Trends Biochem. Sci. 29:579-585. [DOI] [PubMed] [Google Scholar]
  • 28.Goorha, R. 1982. Frog virus 3 DNA replication occurs in two stages. J. Virol. 43:519-528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Goorha, R., G. Murti, A. Granoff, and R. Tirey. 1978. Macromolecular synthesis in cells infected by frog virus 3. VIII. The nucleus is a site of frog virus 3 DNA and RNA synthesis. Virology 84:32-50. [DOI] [PubMed] [Google Scholar]
  • 30.Goorha, R., and K. G. Murti. 1982. The genome of frog virus 3, an animal DNA virus, is circularly permuted and terminally redundant. Proc. Natl. Acad. Sci. USA 79:248-252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Gordon, D., C. Abajian, and P. Green. 1998. Consed: a graphical tool for sequence finishing. Genome Res. 8:195-202. [DOI] [PubMed] [Google Scholar]
  • 32.Guindon, S., and O. Gascuel. 2003. A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst. Biol. 52:696-704. [DOI] [PubMed] [Google Scholar]
  • 33.Hall, D. W., and D. W. Anthony. 1971. Pathology of a mosquito iridescent virus (MIV) infecting Aedes taeniorhynchus. J. Invertebr. Pathol. 18:61-69. [DOI] [PubMed] [Google Scholar]
  • 34.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 genomic analysis of the mandarin fish infectious spleen and kidney necrosis iridovirus. Virology 291:126-139. [DOI] [PubMed] [Google Scholar]
  • 35.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]
  • 36.Hershko, A., and A. Ciechanover. 1998. The ubiquitin system. Annu. Rev. Biochem. 67:425-479. [DOI] [PubMed] [Google Scholar]
  • 37.Home, W. A., S. Tajbakhsh, and V. L. Seligy. 1990. Molecular cloning and characterization of a late Tipula iridescent virus gene. Gene 94:243-248. [DOI] [PubMed] [Google Scholar]
  • 38.Huang, X., and A. Madan. 1999. CAP3: a DNA sequence assembly program. Genome Res. 9:868-877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Huelsenbeck, J. P., and F. Ronquist. 2001. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17:754-755. [DOI] [PubMed] [Google Scholar]
  • 40.Iyer, L. M., E. V. Koonin, and L. Aravind. 2002. Extensive domain shuffling in transcription regulators of DNA viruses and implications for the origin of fungal APSES transcription factors. Genome Biol. 3:RESEARCH 0012.1-0012.11. [Online.] http://genomebiology.com/2002/3/3/RESEARCH/0012. [DOI] [PMC free article] [PubMed]
  • 41.Jakob, N. J., and G. Darai. 2002. Molecular anatomy of Chilo iridescent virus genome and the evolution of viral genes. Virus Genes 25:299-316. [DOI] [PubMed] [Google Scholar]
  • 42.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]
  • 43.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]
  • 44.Jones, D. T., W. R. Taylor, and J. M. Thornton. 1994. A model recognition approach to the prediction of all-helical membrane protein structure and topology. Biochemistry 33:3038-3049. [DOI] [PubMed] [Google Scholar]
  • 45.Kang, W., M. Suzuli, Z. Evgueni, K. Okano, and S. Maeda. 1999. Characterization of baculovirus repeated open reading frames (bro) in Bombyx mori nucleopolyhedrovirus. J. Virol. 73:10339-10345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Keck, J. G., C. J. Baldick, and B. Moss. 1990. Role of DNA replication in vaccinia virus gene expression: a naked template is required for transcription of three late trans-activator genes. Cell 61:801-809. [DOI] [PubMed] [Google Scholar]
  • 47.Kelly, D. C., M. D. Ayres, T. Lescott, J. S. Robertson, and G. M. Happ. 1979. A small iridescent virus (type 29) isolated from Tenebro molitor: a comparison of its proteins and antigens with six other iridescent viruses. J. Gen. Virol. 42:95. [Google Scholar]
  • 48.Krishna, T. S., X.-P. Kong, S. Gary, P. M. Burgers, and J. Kuriyan. 1994. Crystal structure of the eukaryotic DNA polymerase processivity factor PCNA. Cell 79:1233-1243. [DOI] [PubMed] [Google Scholar]
  • 49.Lassmann, T., and E. L. Sonnhammer. 2005. Kalign—an accurate and fast multiple sequence alignment algorithm. BMC Bioinformatics 6:298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Lavrukhin, O. V., J. M. Fortune, T. G. Wood, D. E. Burbank, J. L. Van Etten, N. Osheroff, and S. Lloyd. 2000. Topoisomerase II from Chlorella virus PBCV-1. J. Biol. Chem. 275:6915-6921. [DOI] [PubMed] [Google Scholar]
  • 51.Linley, J. R., and H. T. Nielsen. 1968. Transmission of mosquito iridescent virus in Aedes taeniorhynchus. I. Laboratory experiments. J. Invertebr. Pathol. 12:7-16. [DOI] [PubMed] [Google Scholar]
  • 52.Linley, J. R., and H. T. Nielsen. 1968. Transmission of mosquito iridescent virus in Aedes taeniorhynchus. II. Experiments related to transmission in nature. J. Invertebr. Pathol. 12:17-24. [DOI] [PubMed] [Google Scholar]
  • 53.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 coiodes. Virology 339:81-100. [DOI] [PubMed] [Google Scholar]
  • 54.Marakova, K. S., L. Aravind, and E. V. Kooning. 2000. A novel superfamily of predicted cysteine proteases from eukaryotes, viruses and Chlamydia pneumoniae. Trends Biochem. Sci. 25:50-52. [DOI] [PubMed] [Google Scholar]
  • 55.Marina, C. F., J. I. Arredondo-Jimenez, A. Castillo, and T. Williams. 1999. Sublethal effects of iridovirus disease in a mosquito. Oecologia 119:383-388. [DOI] [PubMed] [Google Scholar]
  • 56.McKune, K., K. L. Richards, A. M. Edwards, R. A. Young, and N. A. Woychik. 1993. Rpb7, one of two dissociable subunits of yeast RNA polymerase II, is essential for cell viability. Yeast 9:295-299. [DOI] [PubMed] [Google Scholar]
  • 57.McLennan, A. G. 1999. The MutT motif family of nucleotide phosphohydrolases in man and human pathogens. Int. J. Mol. Med. 4:79-89. [DOI] [PubMed] [Google Scholar]
  • 58.McMillan, N. A., S. Davison, and J. Kalmakoff. 1990. Comparison of the genomes of two sympatric iridescent viruses (types 9 and 16). Arch. Virol. 114:277-284. [DOI] [PubMed] [Google Scholar]
  • 59.McMillan, N. A., and J. Kalmakoff. 1994. RNA transcript mapping of the Wiseana iridescent virus genome. Virus Res. 32:343-352. [DOI] [PubMed] [Google Scholar]
  • 60.Nachmias, B., Y. Ashhab, and D. Ben-Yehuda. 2004. The inhibitor of apoptosis protein family (IAPs): an emerging therapeutic target in cancer. Semin. Cancer Biol. 14:231-243. [DOI] [PubMed] [Google Scholar]
  • 61.Nakai, K., and P. Horton. 1999. PSORT: a program for detecting sorting signals in proteins and predicting their subcellular localization. Trends Biochem. Sci. 24:34-36. [DOI] [PubMed] [Google Scholar]
  • 62.Parks, W. C., C. L. Wilson, and Y. S. Lopez-Boado. 2004. Matrix metalloproteinases as modulators of inflammation and innate immunity. Nat. Rev. 4:617-629. [DOI] [PubMed] [Google Scholar]
  • 63.Pearson, W. R. 1990. Rapid and sensitive sequence comparison with FASTP and FASTA. Methods Enzymol. 183:63-98. [DOI] [PubMed] [Google Scholar]
  • 64.Salzberg, S. L., A. L. Delcher, S. Kasif, and O. White. 1998. Microbial gene identification using interpolated Markov models. Nucleic Acids Res. 26:544-548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Sanger, F., S. Nicklen, and A. R. Coulson. 1977. DNA sequencing with chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA 74:5463-5467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Schmidt, H. A., K. Strimmer, M. Vingron, and A. von Haeseler. 2002. TREE-PUZZLE: maximum likelihood phylogenetic analysis using quartets and parallel computing. Bioinformatics 18:502-504. [DOI] [PubMed] [Google Scholar]
  • 67.Shors, T., J. G. Keck, and B. Moss. 1999. Down regulation of gene expression by the vaccinia virus D10 protein. J. Virol. 73:791-796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Snijder, E. J., A. L. Wassenaar, W. J. Spaan, and A. E. Gorbalenya. 1995. The arterivirus Nsp2 protease. An unusual cysteine protease with primary structure similarities to both papain-like and chymotrypsin-like proteases. J. Biol. Chem. 270:16671-16676. [DOI] [PubMed] [Google Scholar]
  • 69.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]
  • 70.Sonnhammer, E. L., S. R. Eddy, E. Birney, A. Bateman, and R. Durbin. 1998. Pfam: multiple sequence alignments and HMM-profiles of protein domains. Nucleic Acids Res. 26:320-322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Stasiak, K., M.-V. Demattei, B. A. Federici, and Y. Bigot. 2000. Phylogenetic position of the DpAV-4a ascovirus DNA polymerase among viruses with a large double-stranded DNA genome. J. Gen. Virol. 81:3059-3072. [DOI] [PubMed] [Google Scholar]
  • 72.Subramanian, A. R., J. Weyer-Menkhoff, M. Kaufmann, and B. Morgenstern. 2005. DIALIGN-T: an improved algorithm for segment-based multiple sequence alignment. BMC Bioinformatics 6:66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Tan, W. G. H., T. J. Barkman, V. G. Chinchar, and K. Essani. 2004. Comparative genomic analysis of frog virus 3, type species of the genus Ranavirus (family Iridoviridae). Virology 323:70-84. [DOI] [PubMed] [Google Scholar]
  • 74.Theilmann, D. A., and S. Stewart. 1992. Tandemly repeated sequence at the 3′ end of the IE-2 gene of the baculovirus Orgyia pseudotsugata multicapsid nuclear polyhedrosis virus is an enhancer element. Virology 187:97-106. [DOI] [PubMed] [Google Scholar]
  • 75.Thompson, J. D., D. G. Higgins, and T. J. Gibson. 1994. CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22:4673-4680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Tidona, C. A., and G. Darai. 1997. The complete DNA sequence of lymphocystis disease virus. Virology 230:207-216. [DOI] [PubMed] [Google Scholar]
  • 77.Todone, F., P. Brick, F. Werner, R. O. J. Weinziert, and S. Onesti. 2001. Structure of an archeal homolog of the eukaryotic RNA polymerase II RPB4/RPB7 complex. Mol. Cell 8:1137-1143. [DOI] [PubMed] [Google Scholar]
  • 78.Tsai, C.-T., J.-W. Ting, M.-H. Wu, M.-F. Wu, I.-C. Guo, and C.-Y. Chang. 2005. Complete genomic 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]
  • 79.Ueda, K., A. Yamashita, J. Ishikawa, M. Shimada, T. O. Watsuji, Morimura, H. Ikeda, M. Hattori, and T. Beppu. 2004. Genome sequence of Symbiobacterium thermophilum, an uncultivable bacterium that depends on microbial commensalism. Nucleic Acids Res. 32:4937-4944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Undeen, A. H., and T. Fukuda. 1994. Effects of host resistance and injury on the susceptibility of Aedes taeniorhynchus to mosquito iridescent virus. J. Am. Mosq. Control Assoc. 10:64-66. [PubMed] [Google Scholar]
  • 81.Van Wart, H. E., and H. Birkedal-Hansen. 1990. The cystein switch: a principle of regulation of metalloproteinase activity with potential applicability to the entire matrix metalloproteinase gene family. Proc. Natl. Acad. Sci. USA 87:5578-5582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Wagner, G. W., and J. D. Paschke. 1977. A comparison of the DNA of R and T strains of mosquito iridescent virus. Virology 81:298-308. [DOI] [PubMed] [Google Scholar]
  • 83.Wang, J. W., R. Q. Deng, X. Z. Wang, Y. S. Huang, K. Xing, J. H. Feng, J. G. He, and Q. X. Long. 2003. Cladistic analysis of iridoviruses based on protein and DNA sequences. Arch. Virol. 148:2181-2194. [DOI] [PubMed] [Google Scholar]
  • 84.Webby, R. J., and J. Kalmakoff. 1998. Sequence comparison of the major capsid protein gene from 18 diverse iridoviruses. Arch. Virol. 143:1949-1966. [DOI] [PubMed] [Google Scholar]
  • 85.Wei, W., D. Dorjsuren, Y. Lin, W. Qin, T. Nomura, N. Hayashi, and S. Murakami. 2001. Direct interaction between the subunit RAP30 of transcription factor IIF (TFIIF) and RNA polymerase subunit 5, which contributes to the association between TFIIF and RNA polymerase II. J. Biol. Chem. 276:12266-12273. [DOI] [PubMed] [Google Scholar]
  • 86.Wertz, I. E., K. M. O'Rourke, H. Zhou, M. Eby, L. Aravind, S. Seshagiri, P. Wu, C. Wiesmann, R. Baker, D. L. Boone, A. Ma, E. V. Koonin, and V. M. Dixit. 2004. De-ubiquitination and ubiquitin ligase domains of A20 downregulate NF-κB signalling. Nature 430:694-699. [DOI] [PubMed] [Google Scholar]
  • 87.Wesley, R. D., J. C. Quintero, and C. A. Mebus. 1984. Extraction of viral DNA from erythrocytes of swine with acute African swine fever. Am. J. Vet. Res. 45:1127-1131. [PubMed] [Google Scholar]
  • 88.Williams, T. 1994. Comparative studies of iridoviruses: further support for a new classification. Virus Res. 33:99-121. [DOI] [PubMed] [Google Scholar]
  • 89.Williams, T. 1998. Invertebrate iridescent viruses. Plenum Press, New York, N.Y.
  • 90.Williams, T. 1996. The iridoviruses. Adv. Virus Res. 46:345-412. [DOI] [PubMed] [Google Scholar]
  • 91.Williams, T., and J. S. Cory. 1994. Proposals for a new classification of iridescent viruses. J. Gen. Virol. 75:1291-1301. [DOI] [PubMed] [Google Scholar]
  • 92.Willis, D. B., and A. Granoff. 1985. Trans activation of an immediate-early frog virus 3 promoter by a virion protein. J. Virol. 56:495-501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Woodard, D. B., and H. C. Chapman. 1968. Laboratory studies with the mosquito iridescent virus (MIV). J. Invertebr. Pathol. 11:296-301. [DOI] [PubMed] [Google Scholar]
  • 94.Xeros, N. 1954. A second virus disease of the leather jacket, Tipula paludosa. Nature 174:562. [Google Scholar]
  • 95.Yan, X., N. H. Olson, J. L. Van Etten, M. Bergoin, M. G. Rossmann, and T. S. Baker. 2000. Structure and assembly of large lipid-containing dsDNA viruses. Nat. Struct. Biol. 7:101-103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Yeo, M., P. S. Lin, M. E. Dahmus, and G. N. Gill. 2003. A novel RNA polymerase II C-terminal domain phosphatase that preferentially dephosphorylates serine 5. J. Biol. Chem. 278:26078-26085. [DOI] [PubMed] [Google Scholar]
  • 97.Zemskov, E. A., W. Kang, and S. Maeda. 2000. Evidence for nucleic acid binding ability and nucleosome association of Bombyx mori nucleopolyhedrovirus BRO proteins. J. Virol. 74:6784-6789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Zhang, Q.-Y., F. Xiao, J. Xie, Z.-Q. Li, and J.-F. Gui. 2004. Complete genomic sequence of lymphocystis disease virus isolated from China. J. Virol. 78:6982-6994. [DOI] [PMC free article] [PubMed] [Google Scholar]

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

RESOURCES