TY - JOUR AU - Storfer, Andrew AB - Abstract Ranaviruses, a genus of the Iridoviridae, are large double-stranded DNA viruses that infect cold-blooded vertebrates worldwide. Ranaviruses have caused severe epizootics in commercial frog and fish populations, and are currently classified as notifiable pathogens in international trade. Previous work shows that a ranavirus that infects tiger salamanders throughout Western North America (Ambystoma tigrinum virus, or ATV) is in high prevalence among salamanders in the fishing bait trade. Bait ATV strains have elevated virulence and are transported long distances by humans, providing widespread opportunities for pathogen pollution. We sequenced the genomes of 15 strains of ATV collected from tiger salamanders across western North America and performed phylogenetic and population genomic analyses and tests for recombination. We find that ATV forms a monophyletic clade within the rest of the Ranaviruses and that it likely emerged within the last several thousand years, before human activities influenced its spread. We also identify several genes under strong positive selection, some of which appear to be involved in viral virulence and/or host immune evasion. In addition, we provide support for the pathogen pollution hypothesis with evidence of recombination among ATV strains, and potential bait-endemic strain recombination. Ranavirus, Ambystoma tigrinum virus, range expansion Emerging infectious diseases are increasingly appreciated as a leading health concern for humans, wildlife, and economically important agricultural populations (Daszak et al. 2000; Smith et al. 2009). Indeed, pathogens are now listed as a leading cause of species’ declines and extinctions (De Castro and Bolker 2005; Smith et al. 2006). Ranaviruses, a genus of the Iridoviridae, are globally-distributed pathogens of amphibians, reptiles and commercial fish species (Chinchar 2002; Miller et al. 2011). These large, double-stranded DNA viruses are considered emerging due to increases in incidence and geographic range over the last 30 years (Daszak et al. 2000; Chinchar 2002; Miller et al. 2011). Ranaviruses are now classified as notifiable pathogens in international trade because of their effects on commercial and wildlife populations (Schloegel et al. 2010). Pathogen pollution is of particular concern, whereby non-native ranavirus strains are introduced into host populations with which they have no evolutionary history, potentially leading to large scale epizootics (Cunningham et al. 2003). Phylogenetic analyses provide several lines of evidence for host switching events among the 10 completely sequenced Ranavirus genomes (Jancovich et al. 2010; Abrams et al. 2013). These analyses suggest that TFV (tiger frog virus) and GIV (grouper iridovirus) are likely strains of the geographically-distant FV3 (frog virus 3) and SGIV (Singapore grouper iridovirus), respectively (Chinchar et al. 2011). In addition, strains isolated from different vertebrate classes, such as STIV (soft-shelled turtle iridovirus; Reptilia) and FV3 (Amphibia) are very similar in genome organization, and, like RGV (Lei et al. 2012), both have truncated versions of the viral homolog of eukaryotic translational initiation factor 2α [vIF-2α (Huang et al. 2009; Tan et al. 2004)]. Additional phylogenetic analyses, combined with comparisons of genomic organization, suggest that the most recent common ancestor of all ranaviruses was a strain that infected fish (Jancovich et al. 2010; Chinchar et al. 2011). Evidence of a fish–amphibian host switch comes from the strong collinearity of EHNV (epizootic hematopoietic necrosis virus; isolated from rainbow trout) and ATV (Ambystoma tigrinum virus, isolated from salamanders). A possible mechanism for host switching among vertebrate classes was inferred from evidence of positive selection on 12 ranavirus genes; six of these genes were apparently acquired as a result of host switches (Abrams et al. 2013). The broad host range and propensity for host switching among ranavirus strains has led to widespread concerns surrounding the potential effects of pathogen pollution. Two ranavirus strains that infect amphibians have received considerable attention in this regard. FV3 and FV3-like strains, which have a broad host range, are commonly found when international shipments of bullfrogs are tested using PCR-based methods (Schloegel et al. 2009). In addition, ATV—a ranavirus that infects tiger salamanders across the North American cordillera (Jancovich et al. 1997, 2003, 2005)—has been found to have > 80% prevalence in tiger salamanders in bait shops across Arizona, Colorado, and New Mexico (Picco and Collins 2008). ATV causes seasonally recurrent epidemics, primarily in larval tiger salamander populations that inhabit natural and human-made ponds throughout western North America (Jancovich et al. 1997, 2005; Brunner et al. 2004; Figure 1). Phylogeographic analyses suggest that long-distance ATV dispersal follows highways that are known to be used for transporting bait (Jancovich et al. 2005), and angler surveys confirmed release of unused live bait salamanders into lakes and ponds (Picco and Collins 2008). Thus, extensive opportunities exist for pathogen pollution by ATV. Figure 1 Open in new tabDownload slide Locations of the tiger salamander populations from which the strains of Ambystoma tigrinum virus (ATV) sequenced in this study were obtained. Strains marked in red were collected in a bait shop (ORV) and from a captive axolotl colony (AXO), and hence their location of origin was not known. A reference genome of ATV has been sequenced (Jancovich et al. 2003), and phylogenetic analysis of ATV strains (using DNA sequences comprising about 2% of the genome) throughout the western United States suggests monophyly (Jancovich et al. 2005; Storfer et al. 2007). However, ATV appears to evolve differently in different host populations. That is, of nine genes (ORFs) that have a putative function in host immune evasion and/or viral virulence (Jancovich et al. 2003), two were under positive selection among tiger salamander host populations (Ridenhour and Storfer 2008). Indeed, there is also evidence of coevolution of ATV and tiger salamander hosts as comparative phylogenetic analyses show strong concordance between host and virus phylogenetic trees (Storfer et al. 2007). Thus, it is important to understand genomic variation among ATV strains to help assess the potential effects of pathogen pollution, as coevolutionary dynamics of hosts and pathogens could be disrupted by introduction of non-native bait strains. Here, we sequenced the complete genomes of 15 ATV native and bait-associated strains across a broad geographic area to assess genomic diversity, identify genes that are experiencing strong positive selection, and to test for evidence of recombination. Recombination has been reported to occur among ranaviruses (e.g., Abrams et al. 2013), but has not been verified or quantified. A previous study showed that an ATV strain isolated from a bait shop had significantly greater virulence than endemic, coevolved strains (Storfer et al. 2007). Thus, recombination among non-native and native ATV strains could enhance virulence, leading to severe epizootics and possibly local extirpation. We also conducted phylogenetic analyses of ATV in comparison to other Ranavirus genomic sequences to provide a rigorous test of monophyly vs. a pattern of nonmonophyly resulting from repeated host-switches, as well as timing of the emergence of ATV to assess human involvement in its geographic spread. Materials and Methods Strain collection The genome sequences of 11 previously sequenced Ranavirus strains were obtained from GenBank: Ambystoma tigrinum virus (ATV), Andrias davidianus ranavirus (ADRV), common midwife toad virus (CMTV), tiger frog virus (TFV), frog virus 3 (FV3), Rana grylio virus (RGV), soft-shelled turtle iridovirus (STIV), European sheatfish virus (ESV), epizootic hematopoietic virus (EHNV), grouper iridovirus (GIV), and Singapore grouper iridovirus (SGIV) (Table 1). The genome sequences of 15 additional Ranavirus isolates collected from North America between 1997 and 2003 are described for the first time here (Table 1). Reads and assemblies for these 15 strains have been deposited at NCBI in BioProject PRJNA257291. Locations, collection dates, hosts, and number of reads obtained for strains included in this study Table 1 Locations, collection dates, hosts, and number of reads obtained for strains included in this study Strain . Full Name . Location . Year . Host . Reference . NCBI Accession . Assembly Size . Reads . Previously sequenced  ATV Ambystoma tigrinum virus Santa Cruz, AZ (31°24’ N, 110°27’W) 1995 Ambystoma tigrinum Jancovich et al. 2003 NC_005832.1 106,332  ADRV Andrias davidianus ranavirus Shanxi Province, China 2010 Andrias davidianus Jiang et al. 2011; Wang et al. 2014; Geng et al. 2011 KF033124.1 106,719  TFV Tiger frog virus Nanhai, Guangdong, China ∼1999 Rana tigrina rugulosa He et al. 2002 AF389451.1 105,057  EHNV Epizootic hematopoietic necrosis virus Lake Nillahcootie, Benalla, Australia 1984 Percia fluviatilis Langdon et al. 1986; Jancovich et al. 2010 FJ433873.1 127,011  FV3 Frog virus 3 Wisconsin / Minnesota 1962 Rana pipiens Tan et al. 2004; Granoff and Came 1965 AY548484.1 105,903  CMTV Common midwife toad virus Picos de Europa Nat’l Park, Spain 2007 Alytes obstetricans Balseiro et al. 2009; Mavian et al. 2012a JQ231222.1 106,878  STIV Soft-shelled turtle iridovirus Shenzhen, China 1997 Trionyx sinensis Chen et al. 1999; Huang et al. 2009 EU627010.1 105,890  RGV Rana grylio virus Wuhan, Hubei, China 1995 Rana grylio Zhang et al. 2001; Lei et al. 2012 JQ654586.1 105,791  ESV European sheatfish virus Germany 1989 Silurus glanis Mavian et al. 2012b JQ724856.1 127,732  GIV Grouper iridovirus Taiwan 2000 or earlier Epinephelus awoara Lai et al. 2000; Tsai et al. 2005 AY666015.1 139,793  SGIV Singapore grouper iridovirus Singapore 1998 Epinephelus tauvina Qin et al. 2001; Song et al. 2004 AY521625.1 140,131 Sequenced here  AXO Axlotl colony, Indiana University 2001 Ambystoma mexicanum Jancovich et al. 2005 KR075872 105,504 565,445  CAP Cap Mountain Pond, CO (38°38’N, 105°24’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075886 106,004 543,964  DO-ATV “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075885 105,936 582,711  DOT312 “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2003 Ambystoma tigrinum J. Brunner, unpublished data KR075883 107,829 726,469  GUFFY Guffy Pond, CO (36°37’N, 105°22’W) 2001 Ambystoma tigrinum Jancovich et al. 2005 KR075882 106,437 931,263  HEIDI “Heidi Tank”, AZ (31°20’35’’N, 110°33’51’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075873 106,230 735,634  JMH “Joe’s Mud Hole”, AZ (36°34’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075881 106,380 659,985  NM Paige Well, NM, 33° 06’ N 107° 39’ W 2004 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075880 107,371a 554,072  ORV Phoenix, AZ bait shop 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075874 106,018 1,311,233  SLV Snipe Lake, AZ (36°31’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075878 106,722 487,984  TSMB Boissevain, MB, Canada (49°14’N, 100°10’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075875 106,526 448,595  UTAH Lake Desolation, UT (40°39’N, 111°36’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075877 106,198 625,978  YEL Yellowstone National Park, USA 2003 Ambystoma tigrinum J. Eastman and A. Storfer, unpublished data KR075876 105,922 654,185  DAL1 Dalmeny SK, Canada (52°20’N, 106°45’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075884 105,354 667,658  RRV Regina, SK, Canada (50°30’N, 104°49’W) 1997 Ambystoma tigrinum Jancovich et al. 2005 KR075879 106,971 539,385 Strain . Full Name . Location . Year . Host . Reference . NCBI Accession . Assembly Size . Reads . Previously sequenced  ATV Ambystoma tigrinum virus Santa Cruz, AZ (31°24’ N, 110°27’W) 1995 Ambystoma tigrinum Jancovich et al. 2003 NC_005832.1 106,332  ADRV Andrias davidianus ranavirus Shanxi Province, China 2010 Andrias davidianus Jiang et al. 2011; Wang et al. 2014; Geng et al. 2011 KF033124.1 106,719  TFV Tiger frog virus Nanhai, Guangdong, China ∼1999 Rana tigrina rugulosa He et al. 2002 AF389451.1 105,057  EHNV Epizootic hematopoietic necrosis virus Lake Nillahcootie, Benalla, Australia 1984 Percia fluviatilis Langdon et al. 1986; Jancovich et al. 2010 FJ433873.1 127,011  FV3 Frog virus 3 Wisconsin / Minnesota 1962 Rana pipiens Tan et al. 2004; Granoff and Came 1965 AY548484.1 105,903  CMTV Common midwife toad virus Picos de Europa Nat’l Park, Spain 2007 Alytes obstetricans Balseiro et al. 2009; Mavian et al. 2012a JQ231222.1 106,878  STIV Soft-shelled turtle iridovirus Shenzhen, China 1997 Trionyx sinensis Chen et al. 1999; Huang et al. 2009 EU627010.1 105,890  RGV Rana grylio virus Wuhan, Hubei, China 1995 Rana grylio Zhang et al. 2001; Lei et al. 2012 JQ654586.1 105,791  ESV European sheatfish virus Germany 1989 Silurus glanis Mavian et al. 2012b JQ724856.1 127,732  GIV Grouper iridovirus Taiwan 2000 or earlier Epinephelus awoara Lai et al. 2000; Tsai et al. 2005 AY666015.1 139,793  SGIV Singapore grouper iridovirus Singapore 1998 Epinephelus tauvina Qin et al. 2001; Song et al. 2004 AY521625.1 140,131 Sequenced here  AXO Axlotl colony, Indiana University 2001 Ambystoma mexicanum Jancovich et al. 2005 KR075872 105,504 565,445  CAP Cap Mountain Pond, CO (38°38’N, 105°24’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075886 106,004 543,964  DO-ATV “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075885 105,936 582,711  DOT312 “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2003 Ambystoma tigrinum J. Brunner, unpublished data KR075883 107,829 726,469  GUFFY Guffy Pond, CO (36°37’N, 105°22’W) 2001 Ambystoma tigrinum Jancovich et al. 2005 KR075882 106,437 931,263  HEIDI “Heidi Tank”, AZ (31°20’35’’N, 110°33’51’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075873 106,230 735,634  JMH “Joe’s Mud Hole”, AZ (36°34’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075881 106,380 659,985  NM Paige Well, NM, 33° 06’ N 107° 39’ W 2004 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075880 107,371a 554,072  ORV Phoenix, AZ bait shop 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075874 106,018 1,311,233  SLV Snipe Lake, AZ (36°31’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075878 106,722 487,984  TSMB Boissevain, MB, Canada (49°14’N, 100°10’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075875 106,526 448,595  UTAH Lake Desolation, UT (40°39’N, 111°36’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075877 106,198 625,978  YEL Yellowstone National Park, USA 2003 Ambystoma tigrinum J. Eastman and A. Storfer, unpublished data KR075876 105,922 654,185  DAL1 Dalmeny SK, Canada (52°20’N, 106°45’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075884 105,354 667,658  RRV Regina, SK, Canada (50°30’N, 104°49’W) 1997 Ambystoma tigrinum Jancovich et al. 2005 KR075879 106,971 539,385 a Draft assembly with two contigs; all other assemblies contain only one contig. Open in new tab Table 1 Locations, collection dates, hosts, and number of reads obtained for strains included in this study Strain . Full Name . Location . Year . Host . Reference . NCBI Accession . Assembly Size . Reads . Previously sequenced  ATV Ambystoma tigrinum virus Santa Cruz, AZ (31°24’ N, 110°27’W) 1995 Ambystoma tigrinum Jancovich et al. 2003 NC_005832.1 106,332  ADRV Andrias davidianus ranavirus Shanxi Province, China 2010 Andrias davidianus Jiang et al. 2011; Wang et al. 2014; Geng et al. 2011 KF033124.1 106,719  TFV Tiger frog virus Nanhai, Guangdong, China ∼1999 Rana tigrina rugulosa He et al. 2002 AF389451.1 105,057  EHNV Epizootic hematopoietic necrosis virus Lake Nillahcootie, Benalla, Australia 1984 Percia fluviatilis Langdon et al. 1986; Jancovich et al. 2010 FJ433873.1 127,011  FV3 Frog virus 3 Wisconsin / Minnesota 1962 Rana pipiens Tan et al. 2004; Granoff and Came 1965 AY548484.1 105,903  CMTV Common midwife toad virus Picos de Europa Nat’l Park, Spain 2007 Alytes obstetricans Balseiro et al. 2009; Mavian et al. 2012a JQ231222.1 106,878  STIV Soft-shelled turtle iridovirus Shenzhen, China 1997 Trionyx sinensis Chen et al. 1999; Huang et al. 2009 EU627010.1 105,890  RGV Rana grylio virus Wuhan, Hubei, China 1995 Rana grylio Zhang et al. 2001; Lei et al. 2012 JQ654586.1 105,791  ESV European sheatfish virus Germany 1989 Silurus glanis Mavian et al. 2012b JQ724856.1 127,732  GIV Grouper iridovirus Taiwan 2000 or earlier Epinephelus awoara Lai et al. 2000; Tsai et al. 2005 AY666015.1 139,793  SGIV Singapore grouper iridovirus Singapore 1998 Epinephelus tauvina Qin et al. 2001; Song et al. 2004 AY521625.1 140,131 Sequenced here  AXO Axlotl colony, Indiana University 2001 Ambystoma mexicanum Jancovich et al. 2005 KR075872 105,504 565,445  CAP Cap Mountain Pond, CO (38°38’N, 105°24’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075886 106,004 543,964  DO-ATV “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075885 105,936 582,711  DOT312 “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2003 Ambystoma tigrinum J. Brunner, unpublished data KR075883 107,829 726,469  GUFFY Guffy Pond, CO (36°37’N, 105°22’W) 2001 Ambystoma tigrinum Jancovich et al. 2005 KR075882 106,437 931,263  HEIDI “Heidi Tank”, AZ (31°20’35’’N, 110°33’51’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075873 106,230 735,634  JMH “Joe’s Mud Hole”, AZ (36°34’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075881 106,380 659,985  NM Paige Well, NM, 33° 06’ N 107° 39’ W 2004 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075880 107,371a 554,072  ORV Phoenix, AZ bait shop 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075874 106,018 1,311,233  SLV Snipe Lake, AZ (36°31’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075878 106,722 487,984  TSMB Boissevain, MB, Canada (49°14’N, 100°10’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075875 106,526 448,595  UTAH Lake Desolation, UT (40°39’N, 111°36’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075877 106,198 625,978  YEL Yellowstone National Park, USA 2003 Ambystoma tigrinum J. Eastman and A. Storfer, unpublished data KR075876 105,922 654,185  DAL1 Dalmeny SK, Canada (52°20’N, 106°45’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075884 105,354 667,658  RRV Regina, SK, Canada (50°30’N, 104°49’W) 1997 Ambystoma tigrinum Jancovich et al. 2005 KR075879 106,971 539,385 Strain . Full Name . Location . Year . Host . Reference . NCBI Accession . Assembly Size . Reads . Previously sequenced  ATV Ambystoma tigrinum virus Santa Cruz, AZ (31°24’ N, 110°27’W) 1995 Ambystoma tigrinum Jancovich et al. 2003 NC_005832.1 106,332  ADRV Andrias davidianus ranavirus Shanxi Province, China 2010 Andrias davidianus Jiang et al. 2011; Wang et al. 2014; Geng et al. 2011 KF033124.1 106,719  TFV Tiger frog virus Nanhai, Guangdong, China ∼1999 Rana tigrina rugulosa He et al. 2002 AF389451.1 105,057  EHNV Epizootic hematopoietic necrosis virus Lake Nillahcootie, Benalla, Australia 1984 Percia fluviatilis Langdon et al. 1986; Jancovich et al. 2010 FJ433873.1 127,011  FV3 Frog virus 3 Wisconsin / Minnesota 1962 Rana pipiens Tan et al. 2004; Granoff and Came 1965 AY548484.1 105,903  CMTV Common midwife toad virus Picos de Europa Nat’l Park, Spain 2007 Alytes obstetricans Balseiro et al. 2009; Mavian et al. 2012a JQ231222.1 106,878  STIV Soft-shelled turtle iridovirus Shenzhen, China 1997 Trionyx sinensis Chen et al. 1999; Huang et al. 2009 EU627010.1 105,890  RGV Rana grylio virus Wuhan, Hubei, China 1995 Rana grylio Zhang et al. 2001; Lei et al. 2012 JQ654586.1 105,791  ESV European sheatfish virus Germany 1989 Silurus glanis Mavian et al. 2012b JQ724856.1 127,732  GIV Grouper iridovirus Taiwan 2000 or earlier Epinephelus awoara Lai et al. 2000; Tsai et al. 2005 AY666015.1 139,793  SGIV Singapore grouper iridovirus Singapore 1998 Epinephelus tauvina Qin et al. 2001; Song et al. 2004 AY521625.1 140,131 Sequenced here  AXO Axlotl colony, Indiana University 2001 Ambystoma mexicanum Jancovich et al. 2005 KR075872 105,504 565,445  CAP Cap Mountain Pond, CO (38°38’N, 105°24’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075886 106,004 543,964  DO-ATV “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075885 105,936 582,711  DOT312 “Doughnut Tank”, AZ (36°25’20’’N, 105°12’40’’W) 2003 Ambystoma tigrinum J. Brunner, unpublished data KR075883 107,829 726,469  GUFFY Guffy Pond, CO (36°37’N, 105°22’W) 2001 Ambystoma tigrinum Jancovich et al. 2005 KR075882 106,437 931,263  HEIDI “Heidi Tank”, AZ (31°20’35’’N, 110°33’51’’W) 2000 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075873 106,230 735,634  JMH “Joe’s Mud Hole”, AZ (36°34’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075881 106,380 659,985  NM Paige Well, NM, 33° 06’ N 107° 39’ W 2004 Ambystoma tigrinum Ridenhour and Storfer 2008 KR075880 107,371a 554,072  ORV Phoenix, AZ bait shop 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075874 106,018 1,311,233  SLV Snipe Lake, AZ (36°31’N, 112°12’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075878 106,722 487,984  TSMB Boissevain, MB, Canada (49°14’N, 100°10’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075875 106,526 448,595  UTAH Lake Desolation, UT (40°39’N, 111°36’W) 1998 Ambystoma tigrinum Jancovich et al. 2005 KR075877 106,198 625,978  YEL Yellowstone National Park, USA 2003 Ambystoma tigrinum J. Eastman and A. Storfer, unpublished data KR075876 105,922 654,185  DAL1 Dalmeny SK, Canada (52°20’N, 106°45’W) 2000 Ambystoma tigrinum Jancovich et al. 2005 KR075884 105,354 667,658  RRV Regina, SK, Canada (50°30’N, 104°49’W) 1997 Ambystoma tigrinum Jancovich et al. 2005 KR075879 106,971 539,385 a Draft assembly with two contigs; all other assemblies contain only one contig. Open in new tab DNA extraction ATV strains were isolated from infected tiger salamanders at each of the 15 localities by first freezing the entire salamander at –70°. Then, each salamander was homogenized in Eagle’s minimum essential medium (MEM), and the slurry was filtered in preparation for viral cell culture. Cell culture protocols for replication of ATV have been standardized (Jancovich et al. 1997, 2001); the 15 virus isolates to be used herein all produced cytopathology in cultured Epithelioma papilloma cyprini fish cells (Fijan et al. 1983) within 48–72 hr. Virus strains were triple plaque-purified and cultivated with no more than three total passages and equal number of passages per strain, because serial passages through clonal cell lines could result in viral evolution, such as adaptation to the cell line and consequent attenuation in the host (see Ebert 1999). Each strain was replicated to approximately 100 ml in volume in MEM, and viral DNA extraction followed a modified sucrose-gradient protocol (Thurber et al. 2009). In brief, DNA was extracted by lysing the host cells with three freeze–thaw cycles, then incubating with DNase, layering the DNA onto a 25% sucrose pad, and spinning at 16,000 rpm for 8 hr in an ultracentrifuge. The DNA was then purified by treatment with proteinase K and RNase followed by a phenol extraction. Sequencing, assembly, and annotation Libraries were prepared using standard protocols and ATV strains were sequenced at the University of Idaho Genomics Resources Core using 300-bp paired-end reads on an Illumina MiSeq, resulting in a mean of 670,000 reads per strain (Table 1). The reads were cleaned with seqyclean (available from https://github.com/ibest/seqyclean) using ESTs from Pimephales promelas as contaminants, then overlapping read pairs were merged using Flash (v1.2.8; Magoč and Salzberg 2011). A reference-guided assembly was performed using ARC (Hunter et al. 2015), with both ATV and FV3 as the reference genome in separate analyses, followed by manual finishing. For most strains, the best assembly (most continuous) was obtained using SPAdes (v3.0.0; Bankevich et al. 2012) as the assembler, except for ORV and JMH (see Table 1), for which we used Newbler (Marguiles et al. 2005). We obtained assistance with the ORV and JMH assemblies from the University of Idaho Genomics Core. To correct errors in the assembly and remove any reference bias, we aligned the reads for each strain back to its assembly. Any position not covered by ≥ 10 uniquely aligning reads, and with a single base call supported by at least 80% of the uniquely aligning reads with a quality score ≥ 30, and with support on reads aligning to both strands was changed to “N.” If the aligned reads supported a different base than the assembly, the assembly was changed. We relaxed the requirement for alignment to both strands for the first and last 100 bp of the assembly, and we trimmed any remaining terminal ambiguities. The final fold coverage of the assemblies ranged from approximately 350 to 2300. Although relatively long, paired-end reads and very high coverage should result in high accuracy, we verified the assembled sequence at seven loci, including two repetitive regions, using 50,833 bp of Sanger sequence data from 14 strains (no Sanger data were available for DOT312) (Genbank accessions EU512250-EU512270 and EU512376-EU512396; Ridenhour and Storfer 2008). These data included two variable repeat regions. We annotated the genomes using the Rapid Annotation Transfer Tool (RATT; Otto et al. 2011), with ATV, FV3, and CMTV as the reference genomes, ATG as the start codon, TGA, TAA, and TAG as stop codons, no splice site correction, and no pseudogene correction. We used a custom python script to remove potentially redundant annotations (we picked the longest ORF for annotations with the same stop codon). We also removed 20–30 ORFs that were split into multiple pieces, because none of the amphibian-like ranavirus (ALRV) sequences from GenBank have introns or split features. BLASTP (v2.2.28; Camacho et al. 2009) was used to search Swissprot (downloaded April 9, 2014; Boeckmann et al. 2003) to find more descriptive annotations for the genes. Sets of orthologous genes were identified using blastclust (Camacho et al. 2009) using a threshold of 80% amino acid identity across 80% of the length of both sequences; to construct the genealogy with GIV strains, we required only 60% identity across 60% of the length of both sequences. The protein sequences of orthologous genes were aligned using muscle (Edgar 2004), and the protein sequence alignment was then used to construct the nucleotide alignment. We also aligned the nucleotide sequences separately from the protein sequences and considered indels that were not a multiple of 3 bp long to be potential frameshifts. To avoid frameshift mutations being erroneously treated as nonsynonymous mutations, alignments in genes with potential frameshifts were trimmed to the last codon before the frameshift before running phylogenetic or population genomic analyses. Dotplots Pairwise genome comparisons were conducted using the dnadiff program from MUMmer (Kurtz et al. 2004), and dotplots were drawn using custom R software (R Core Team 2014) code. Dotplots are used for visualizing structural differences between pairs of genomes: the axes are position in each genome, and a dot is placed at every location where the two genomes have similar sequence. Identical genomes appear as a diagonal line, while inversions appear as changes in the orientation of matches. Genealogical analyses Genealogical analyses were conducted on the concatenated nucleotide sequences of the 52 genes that were found in all ALRV strains using BEAST (v1.8.0; Drummond et al. 2012) and MrBayes (v3.2.2; Ronquist et al. 2012). Based on model-selection tests conducted with jModelTest (v2.1.4; Darriba et al. 2012) to determine the most likely model of molecular evolution, we used the general time-reversible model with a gamma shape parameter (GTR + G) and four site categories. In BEAST we used a lognormal relaxed clock, empirical base frequencies, and a Bayesian skyline coalescent tree prior, on data partitioned by codon position. The tip date option was used to provide dating information; the tip dates were entered as shown in Table 1 with a precision (accuracy of collection date) of 0.5, except for EHNV, ESV, and FV3, for which we entered a precision of 2.0, and TFV for which we entered a precision of 5.0. BEAST coestimates the divergence times, substitution rates, and genealogy given the sequence data and tip dates and produces estimates of divergence times integrated over the posterior distribution of topologies (Drummond et al. 2006). The tip dates calibrate the divergence and rate estimates, so that the result is in absolute, rather than relative, time. Because we used a relaxed clock, we do not assume that evolutionary rates are constant across all branches; rather the rates are drawn from a lognormal distribution. Nearly all of the priors were left as default, except for the relative rate parameters for the codon positions, which were set to a finite uniform prior. Five independent runs with 200,000,000 iterations each were conducted and combined, after confirming that the results of each run were very similar, with logcombiner (Drummond et al. 2012). The first 10% of each run was discarded as burn-in. We confirmed that the chains converged and had sufficient sampling using tracer (Drummond et al. 2012). Finally, we constructed a maximum clade credible tree (MCCT) using treeannotator (Drummond et al. 2012) with median branch heights. A strict molecular clock was rejected by examining the distribution of the ucld.stdev parameter—the distribution did not overlap with zero. We used BEAST to construct the Bayesian skyline plot for just the ATV strains using the same methods, except that the 71 genes found in all ATV strains were included. We also conducted two runs of MrBayes: one with the same 52 genes used for the BEAST coalescent tree analysis and another with 17 genes found in all ALRV strains plus GIV and SGIV. We used the GTR + G model of evolution with four rate categories and partitioning by codon position. MrBayes was for 12,000,000 iterations, and we verified that the standard deviation of split frequencies was < 0.01. PAML analyses We searched for evidence of positive selection using PAML (v4.7; Yang 2007), which identifies genes with elevated ratios of nonsynonymous to synonymous substitutions (ω). We ran three analyses on the ATV clade (other strains were excluded): site tests comparing a model that included neutral sites and sites under purifying selection (model 8A; specified by setting NSsites = 8, fix_omega = 1, omega = 1) to a model that also allowed positive selection (model 8; NSsites = 8, fix_omega = 0); and two sets of branch-site tests comparing a model with neutral sites and purifying selection at all branches (model 2A; NSsites = 2, model = 2, fix_omega = 1, and omega = 1) to a model that also allows positive selection on particular branches and sites (model 2; NSsites = 2, model = 2, and fix_omega = 0). The branch leading to DAL1 was tested because it appears to be geographically distant from its closest relatives (Figure 1 and Figure 2), and the branches leading to UTAH, ORV, and AXO were tested because these are human-associated or bait strains (Ridenhour and Storfer 2008; J. Eastman and A. Storfer, unpublished data). To identify genes with strong positive selection, we conducted a likelihood ratio test with 1 degree of freedom and divided the p-value by two, as recommended in the PAML manual. In all cases, we set PAML to remove sites with indels or ambiguities, used the branch lengths in the MrBayes genealogy as the starting point, did not assume a molecular clock, and estimated codon frequencies from base frequencies. Only the 93 ORF clusters with ≥ 11 ATV strains and ≥ 15 codons were included. To avoid problems with lack of convergence, we ran both the site and branch-site models three times with different initial values, and chose the runs with the least negative log-likelihood. ORF cluster 94 was removed from the site test results because PAML indicated that every codon was under positive selection in spite of having only two variable sites. Figure 2 Open in new tabDownload slide Bayesian genealogy of Ranavirus. (A) The genealogy constructed with MrBayes using the 17 genes found in all Ranavirus; for the sake of clarity, Bayesian posterior support values (edge labels) are shown only for the older nodes. (B) Relationships constructed using 52 genes found in all amphibian-like ranavirus (ALRV) strains using MrBayes. (A) was rooted using Singapore grouper iridovirus (SGIV) and grouper iridovirus (GIV) as a outgroup. The large, black arrow in (B) shows the location of this outgroup; due to a polytomy in (A), the exact location of the root branch in (B) is ambiguous, although the relative ancestry of the ATV and European sheatfish virus/epizootic hematopoietic virus (ESV/EHNV) clades are not affected by the ambiguity. A genealogy of the ALRV strains constructed using BEAST had an identical unrooted topology and qualitatively similar branch lengths and support values, but with a different root placement in the maximum clade credible tree (shown with a light gray arrow). The closed circle, open circle, and triangle indicate an ATV-like, common midwife toad virus (CMTV)-like, or frog virus 3 (FV3)-like genome structure, respectively. Host taxa are indicated to the right of the tree on (A). Population genetic analysis To characterize the nucleotide diversity of ATV, we calculated Watterson’s θ (θW; Watterson 1975), π (θπ; Kimura 1968) Tajima’s D (DT; Tajima 1989), and the McDonald-Kreitman test statistics (MK test; McDonald and Kreitman 1991) using custom C++ programs based on libsequence (Thornton 2003). We include only biallelic sites without ambiguities. Watterson’s θ and π are coalescent-based estimators of the population mutation rate (θ = 2Neμ) assuming standard neutral conditions and infinite sites or infinite alleles, respectively, that are useful as a sample-size adjusted measurement of nucleotide diversity. Under the infinite alleles and infinite sites models, it is expected that all polymorphisms are biallelic: overall, 95% of polymorphic sites have two alleles, and if five ORF clusters (clusters 12, 66, 82, 86, and 95) that were challenging to align are excluded, > 99% of polymorphic sites have two alleles. Thus, there is no evidence for a serious violation of the infinite sites or alleles models. Tajima’s D summarizes the allele frequency spectrum, and can serve as an indicator of demographic events that result in changes in population size or selective sweeps. Tajima’s D is negative after recovery from a selective sweep or during rapid population expansion (negative D values can also result from background purifying selection), positive under balancing selection or population decline, and 0 under the standard neutral model. We used only genes with data from ≥ 11 of the ATV strains; Tajima’s D was reported only if there were ≥ 3 SNPs in a gene. The MK test compares the ratio of nonsynonymous polymorphism: synonymous polymorphism to the ratio of nonsynonymous divergence: synonymous divergence; excess levels of divergence are indicative of positive selection. For the MK test, we used the ESV/EHNV clade as an outgroup. A modified version of the MK test has been suggested for viral datasets with large numbers of multi-allelic sites (Bhatt et al. 2010); however, as discussed above, our dataset contained few multi-allelic sites. Recombination analysis We tested for the presence of recombination in the ATV clade using the software packages RDP4 (v4.36; Martin et al. 2010) and Rbrothers (v0.0.1; Irvahn et al. 2013). The tests in RDP4 search an alignment for positions where there is a change in relative sequence similarity among strains; RDP4 then attempts to combine the results from multiple methods into a single set of results. Rbrothers searches for changes in phylogeny along an alignment, but does not attempt to identify specific events; rather, it estimates the most likely number of recombination events. We first aligned the ATV clade genomes using mugsy (Angiuoli and Salzberg 2011), and then stitched the aligned blocks together using ATV as a reference, dropping alignment blocks that did not contain ATV (these regions were small, contained few strains, and tended to be poorly aligned). For RDP4, we use the RDP (Martin and Rybicki 2000), Geneconv (Padidam et al. 1999), BootScan (Martin et al. 2005), MaxChi (Smith 1992), Chimaera (Posada and Crandall 2001), SiSscan (Gibbs et al. 2000), and 3Seq (Boni et al. 2007) methods with the default settings during both the primary scan (identification of recombination) and secondary scan (verification) phases. Although these methods differ in algorithmic and statistical details, all search an alignment for positions where a change in relative sequence similarity occurs, using three or four sequences at a time. The defaults include a p-value threshold of 0.05 and a Bonferroni correction for multiple comparisons; due to the size of the dataset, we did not select the option to require phylogenetic evidence for recombination events. The RDP4 package attempts to group similar signals of recombination into single events that may represent ancestral recombination events, and also attempts to identify the source (“minor parent”) of the recombined region. For Rbrothers, which uses a dual multiple change-point model (Minin et al. 2005) to identify positions where the phylogeny constructed using all sequences changes, we used a 4000-bp sliding window (2000-bp step) to generate the candidate trees, and ran the chain for 41,000,000 iterations, including 1,000,000 iterations of burn-in. As recommended by Irvahn et al. (2013), the analysis was repeated for several “lambda_prior” (expected number of breakpoints) values: 1, 5, 10, and 20. Rbrothers does not attempt to characterize the details of recombination events, but rather tries to determine the most likely number of events. To estimate the population scaled recombination rate ρ = 2Ner (where Ne = effective population size and r = recombination rate), we used the interval program in the LDHat package (v2.2; Auton and McVean 2007). The interval program identifies recombination by fitting a coalescent model with recombination and mutation to the sequence data using a Markov Chain Monte Carlo approach. We used the same whole genome alignment as described above and ran interval for 10,000,000 iterations, sampling every 2000 after discarding the first 100,000 iterations as burn-in. Both crossing-over mode and gene-conversion mode were used, with gene conversion tract lengths 25, 50, 100, 500, and 1000; and with block penalties 0, 5, 10, 20, 35, and 50. Before running interval, we removed sites that were present in less than 80% of the strains (approximately 1% of the alignment) or with a minor allele frequency < 0.2. Data availability All reads, assemblies, and annotations generated in this study have been deposited in GenBank under BioProject PRJNA257291. Results Genome size and structure The lengths of the reference-guided genome assemblies ranged from 105,354 to 107,371 bp (Table 1), very similar to the reference sequence, which is 106,332 bp long (Jancovich et al. 2003). The mean depth of coverage was at least 350 for all strains, which is far in excess of the coverage necessary to obtain accurate base calls. In addition, 50,833 bp of Sanger sequencing data from seven regions, including two repetitive regions, revealed only 20 single base errors (assuming all mismatches between Sanger sequencing and assemblies are errors in the assemblies)—an error rate of only 0.04%—and no structural errors. Because Sanger sequencing has errors too, and the sequencing data were obtained 10 years ago, the true error rate may be less than 0.04%. The number of putative ORFs per genome ranged from 95 to 110. There was little variation in gene content with ≤ 3 unique ORFs per strain, and 75% of the genes found in the ATV reference genome (Jancovich et al. 2003) were also found in all the 15 strains sequenced here (see Supporting Information, File S1, Table S2), resulting in 71 ORF clusters in the ATV core genome. The core genome of all the ALRVs (all the strains included in this study except GIV and SGIV) comprised 52 ORF clusters, similar to previous results from Jancovich et al. (2010). Also consistent with previous studies, we found that CMTV (common midwife toad virus; Mavian et al. 2012a) has an inversion relative to ATV; and FV3, RGV (Rana grylio virus), TFV, and STIV have an additional inversion (Figure 3) (Stöhr et al. 2015). However, the other genomes were collinear with ATV, and there was no evidence for large-scale inversions or large indels. The ATV strains were also very similar in nucleotide sequence: genome-wide pairwise nucleotide identity ranged from 98.3 to > 99.9%, with a mean of 99.1%. Pairwise identity between ATV strains and other strains ranged from 95.4 to 97.0%, with a mean of 96.2%. Despite the similarity in gene content, there were three genes present in all ATV strains, but absent from all other strains: ORF clusters 84, 85, and 88. Interestingly, ORF cluster 84, a potential serum response factor-binding protein, shows evidence of positive selection (see below); the functions of clusters 85 and 88 are not known and they did not show significant evidence for positive selection by any of our tests. Figure 3 Open in new tabDownload slide Comparison of the three genome structures observed in the amphibian-like Ranavirus strains. (A) Synteny plots comparing ATV and FV3, and ATV and CMTV. (B) The same alignments represented as dotplots, with ATV on the x-axis. Compared to ATV, CMTV has an inversion (60–90 kb in ATV), and FV3 has an additional inversion encompassing most of the genome. Genealogical relationships We constructed a Bayesian genealogy for ALRV using BEAST (Drummond et al. 2012) and MrBayes (Ronquist et al. 2012) with the concatenated sequences of the core genome. Regardless of the set of genes or software used, strains of ATV formed a well-supported monophyletic clade (Figure 2). Using the dates strains were collected, the age of the ATV clade was estimated at 545 years (20–5672 years is the 95% highest posterior density range) and the age of the root of the tree—the common ancestor of all ALRV strains sampled here—was estimated at 1526 years (53–16,524). To see whether a signal of recent population expansion in ATV or ALRV could be detected, we constructed a Bayesian skyline plot (Drummond et al. 2005) for the ALRV group using BEAST. Effective population sizes over the last several hundred years are stable relative to the size of the confidence intervals, although the point estimate shows a decline over the last 100 years (Figure 4). A plot constructed for just the ATV group also showed a stable effective population size (Figure S1). Figure 4 Open in new tabDownload slide Bayesian skyline plot of the Amphibian-like Ranaviruses. The y-axis is proportional to effective population size and the x-axis is time from collection of the samples (around the year 2000 for most strains). The median estimate is given by the black line and the 95% confidence intervals are shaded in blue. Population genetic diversity To characterize the nucleotide diversity in ATV, we calculated population genetic statistics for the 93 ORF clusters found within at least 11 strains in the ATV clade (Figure 5). On average, across all clusters, θW, the average number of segregating sites per bp, and θπ, the average number of pairwise differences per bp, were both 0.01. Tajima’s D (calculated for 85 genes that had ≥ 3 SNPs) was skewed negative but had a mean near 0; the 95% confidence interval obtained by bootstrapping over genes 200 times was –0.47 to –0.15. Figure 5 Open in new tabDownload slide Distributions and means for population genetic statistics. Statistics were calculated for the ATV clade on genes with at least 11 strains present. θW and θπ are measures of nucleotide diversity. Tajima’s D (DT) describes the shape of the allele frequency spectrum: the neutral expectation is 0, while selective sweeps or population expansion result in negative values and balancing selection or population contraction result in positive values. Selection We searched for ORF clusters under strong positive selection within the ATV clade by testing for elevated ratios of nonsynonymous to synonymous mutations (ω) with PAML (Yang 2007) site and branch-site tests. We also tested for genes with evidence of adaptive divergence between the ATV and ESV/EHNV clades with the McDonald-Kreitman (MK) test. Eight ORF clusters (Table 2) showed significant evidence of positive selection, after a sequential Bonferroni correction for multiple comparisons (eight clusters by site tests, and two clusters by branch-site tests). Sixteen additional clusters also showed a significant signature of positive selection (P < 0.05) if no multiple test correction was performed (Table 2). Several of these genes are potentially involved in interactions with the host immune system or replication machinery; for example, ORF cluster 84 may be a cell cycle regulator. However, some of the clusters with the strongest evidence of positive selection had alignments with many indels, which makes alignment challenging. Two more gene clusters, 93 (3-β-hydroxy-D 5-C27-steroid oxidoreductase-like protein) and 95 (no confident annotation, but shows homology to collagen-like protein 1), showed strong evidence of positive selection in PAML site tests (P ≤ 0.0001). ORF cluster 95 may make up part of the viral surface such as envelope proteins; and ORF cluster 93 is involved in suppression of host immunity (Chinchar 2002; Miller et al. 2011). However, both of these genes included several sequences with mutated start codons or frameshift mutations that likely result in a nonfunctional product. After removing these sequences, cluster 93 no longer showed evidence for positive selection. On the other hand, cluster 95, while now having fewer than our 11-strain cutoff for running PAML tests, still showed evidence for selection by site tests (P < 0.0001) and branch-site tests (P = 0.0001). Overall, the PAML analysis indicated that most nonsynonymous sites in most genes are neutral or under purifying selection in the ATV clade: ω << 1 for most genes (Figure S2). Genes with evidence for positive selection in Ambystoma tigrinum virus (ATV) (omega for positively selected codons, p-value). Table 2 Genes with evidence for positive selection in Ambystoma tigrinum virus (ATV) (omega for positively selected codons, p-value). Gene Cluster . ATV Gene . Site Test . Branch-Site DAL1 . Branch-Site Human-Associated . MK Test . Annotation . 1 ATV_AAP33221.1 0.04, 2.0 DNA-dependent RNA polymerase, subunit a 2 ATV_AAP33249.1 0.05, 2.4 Unknown 6 ATV_AAP33184.1 0.007, 11.2 NTPase 8 ATV_AAP33245.1 0.02 Ribonucleoside diphosphate reductase 9 ATV_AAP33230.1 0.01 Myristylated membrane protein 15 ATV_AAP33229.1 0.002, 11.6 Helicase 16 ATV_AAP33190.1 0.01 Immediate early protein (ICP-46) 23 ATV_AAP33268.1 0.03 Unknown 33 ATV_AAP33200.1 0.005, 42.4 Immediate early protein (ICP-18) 38 ATV_AAP33188.1 0.05 Unknown 40 ATV_AAP33254.1 0.009, 6.8 Unknown 54 ATV_AAP33207.1 0.005, 5.2 Unknown 62 ATV_AAP33247.1 < 0.0001a, 11.0 0.04 Unknown 63 — 0.02, 68.9 Unknown 65 ATV_AAP33212.1 0.0003a, 22.2 Unknown protein with homology to transmembrane proteins 66 ATV_AAP33240.1 < 0.0001a, 16.8 0.004; 81.1 Neurofilament triplet 1H-like protein 67 ATV_AAP33256.1 < 0.0001a, 5.6 0.02, 30.7 Unknown protein with homology to SAP domain-containing protein 73 — 0.02, 7.1 Unknown 75 ATV_AAP33189.1 0.0005a, 17.8 0.002, 1.0 Unknown 82 ATV_AAP33252.1 0.002, 93 0.001 Neurofilament triplet H1-like protein 84 ATV_AAP33250.1 < 0.0001a, 9.3 < 0.0001a, NA < 0.0001a, 218 Unknown protein with homology to serum response factor-binding protein 1 89 ATV_AAP33261.1 0.03, 13.6 Unknown protein with homology to serine/threonine kinase protein 90 — 0.0001a, 16.1 Unknown 95b ATV_AAP33253.1 <0.0001a, 4.3 0.0001a, NA Unknown protein with homology to collagen-like protein 1 Gene Cluster . ATV Gene . Site Test . Branch-Site DAL1 . Branch-Site Human-Associated . MK Test . Annotation . 1 ATV_AAP33221.1 0.04, 2.0 DNA-dependent RNA polymerase, subunit a 2 ATV_AAP33249.1 0.05, 2.4 Unknown 6 ATV_AAP33184.1 0.007, 11.2 NTPase 8 ATV_AAP33245.1 0.02 Ribonucleoside diphosphate reductase 9 ATV_AAP33230.1 0.01 Myristylated membrane protein 15 ATV_AAP33229.1 0.002, 11.6 Helicase 16 ATV_AAP33190.1 0.01 Immediate early protein (ICP-46) 23 ATV_AAP33268.1 0.03 Unknown 33 ATV_AAP33200.1 0.005, 42.4 Immediate early protein (ICP-18) 38 ATV_AAP33188.1 0.05 Unknown 40 ATV_AAP33254.1 0.009, 6.8 Unknown 54 ATV_AAP33207.1 0.005, 5.2 Unknown 62 ATV_AAP33247.1 < 0.0001a, 11.0 0.04 Unknown 63 — 0.02, 68.9 Unknown 65 ATV_AAP33212.1 0.0003a, 22.2 Unknown protein with homology to transmembrane proteins 66 ATV_AAP33240.1 < 0.0001a, 16.8 0.004; 81.1 Neurofilament triplet 1H-like protein 67 ATV_AAP33256.1 < 0.0001a, 5.6 0.02, 30.7 Unknown protein with homology to SAP domain-containing protein 73 — 0.02, 7.1 Unknown 75 ATV_AAP33189.1 0.0005a, 17.8 0.002, 1.0 Unknown 82 ATV_AAP33252.1 0.002, 93 0.001 Neurofilament triplet H1-like protein 84 ATV_AAP33250.1 < 0.0001a, 9.3 < 0.0001a, NA < 0.0001a, 218 Unknown protein with homology to serum response factor-binding protein 1 89 ATV_AAP33261.1 0.03, 13.6 Unknown protein with homology to serine/threonine kinase protein 90 — 0.0001a, 16.1 Unknown 95b ATV_AAP33253.1 <0.0001a, 4.3 0.0001a, NA Unknown protein with homology to collagen-like protein 1 Only significant (P < 0.05) results are shown. In some cases PAML listed the omega value as 999, this indicates that the estimate of Ks was zero; we marked the omega for these genes as NA. The ESV/EHNV clade was used as the outgroup for the MK test. a Significant (P < 0.05) after sequential Bonferroni correction. b Cluster 95 contains fewer than 11 strains after removal of nonfunctional sequences. Open in new tab Table 2 Genes with evidence for positive selection in Ambystoma tigrinum virus (ATV) (omega for positively selected codons, p-value). Gene Cluster . ATV Gene . Site Test . Branch-Site DAL1 . Branch-Site Human-Associated . MK Test . Annotation . 1 ATV_AAP33221.1 0.04, 2.0 DNA-dependent RNA polymerase, subunit a 2 ATV_AAP33249.1 0.05, 2.4 Unknown 6 ATV_AAP33184.1 0.007, 11.2 NTPase 8 ATV_AAP33245.1 0.02 Ribonucleoside diphosphate reductase 9 ATV_AAP33230.1 0.01 Myristylated membrane protein 15 ATV_AAP33229.1 0.002, 11.6 Helicase 16 ATV_AAP33190.1 0.01 Immediate early protein (ICP-46) 23 ATV_AAP33268.1 0.03 Unknown 33 ATV_AAP33200.1 0.005, 42.4 Immediate early protein (ICP-18) 38 ATV_AAP33188.1 0.05 Unknown 40 ATV_AAP33254.1 0.009, 6.8 Unknown 54 ATV_AAP33207.1 0.005, 5.2 Unknown 62 ATV_AAP33247.1 < 0.0001a, 11.0 0.04 Unknown 63 — 0.02, 68.9 Unknown 65 ATV_AAP33212.1 0.0003a, 22.2 Unknown protein with homology to transmembrane proteins 66 ATV_AAP33240.1 < 0.0001a, 16.8 0.004; 81.1 Neurofilament triplet 1H-like protein 67 ATV_AAP33256.1 < 0.0001a, 5.6 0.02, 30.7 Unknown protein with homology to SAP domain-containing protein 73 — 0.02, 7.1 Unknown 75 ATV_AAP33189.1 0.0005a, 17.8 0.002, 1.0 Unknown 82 ATV_AAP33252.1 0.002, 93 0.001 Neurofilament triplet H1-like protein 84 ATV_AAP33250.1 < 0.0001a, 9.3 < 0.0001a, NA < 0.0001a, 218 Unknown protein with homology to serum response factor-binding protein 1 89 ATV_AAP33261.1 0.03, 13.6 Unknown protein with homology to serine/threonine kinase protein 90 — 0.0001a, 16.1 Unknown 95b ATV_AAP33253.1 <0.0001a, 4.3 0.0001a, NA Unknown protein with homology to collagen-like protein 1 Gene Cluster . ATV Gene . Site Test . Branch-Site DAL1 . Branch-Site Human-Associated . MK Test . Annotation . 1 ATV_AAP33221.1 0.04, 2.0 DNA-dependent RNA polymerase, subunit a 2 ATV_AAP33249.1 0.05, 2.4 Unknown 6 ATV_AAP33184.1 0.007, 11.2 NTPase 8 ATV_AAP33245.1 0.02 Ribonucleoside diphosphate reductase 9 ATV_AAP33230.1 0.01 Myristylated membrane protein 15 ATV_AAP33229.1 0.002, 11.6 Helicase 16 ATV_AAP33190.1 0.01 Immediate early protein (ICP-46) 23 ATV_AAP33268.1 0.03 Unknown 33 ATV_AAP33200.1 0.005, 42.4 Immediate early protein (ICP-18) 38 ATV_AAP33188.1 0.05 Unknown 40 ATV_AAP33254.1 0.009, 6.8 Unknown 54 ATV_AAP33207.1 0.005, 5.2 Unknown 62 ATV_AAP33247.1 < 0.0001a, 11.0 0.04 Unknown 63 — 0.02, 68.9 Unknown 65 ATV_AAP33212.1 0.0003a, 22.2 Unknown protein with homology to transmembrane proteins 66 ATV_AAP33240.1 < 0.0001a, 16.8 0.004; 81.1 Neurofilament triplet 1H-like protein 67 ATV_AAP33256.1 < 0.0001a, 5.6 0.02, 30.7 Unknown protein with homology to SAP domain-containing protein 73 — 0.02, 7.1 Unknown 75 ATV_AAP33189.1 0.0005a, 17.8 0.002, 1.0 Unknown 82 ATV_AAP33252.1 0.002, 93 0.001 Neurofilament triplet H1-like protein 84 ATV_AAP33250.1 < 0.0001a, 9.3 < 0.0001a, NA < 0.0001a, 218 Unknown protein with homology to serum response factor-binding protein 1 89 ATV_AAP33261.1 0.03, 13.6 Unknown protein with homology to serine/threonine kinase protein 90 — 0.0001a, 16.1 Unknown 95b ATV_AAP33253.1 <0.0001a, 4.3 0.0001a, NA Unknown protein with homology to collagen-like protein 1 Only significant (P < 0.05) results are shown. In some cases PAML listed the omega value as 999, this indicates that the estimate of Ks was zero; we marked the omega for these genes as NA. The ESV/EHNV clade was used as the outgroup for the MK test. a Significant (P < 0.05) after sequential Bonferroni correction. b Cluster 95 contains fewer than 11 strains after removal of nonfunctional sequences. Open in new tab Recombination An exploratory analysis of recombination, consisting of seven tests (RDP, Geneconv, Recscan/Bootscan, SiSscan, Maxchi, Chimaera, and 3Seq) implemented in the RDP4 package (Martin et al. 2010), a method implemented in Rbrothers (Irvahn et al. 2013), and a method implemented in the interval program of the LDHat package (Auton and McVean 2007), all support evidence of recombination within the ATV clade. RDP4 detected 74 events total (best-supported 20% of events are in Table S1), while the number of recombination events detected by Rbrothers varied by parameter settings (Table 3). Using LDHat [57], our point estimates for the population scaled recombination rate (ρ = 2Ner) ranged from 0.00031 to 0.00033 per nucleotide, depending on parameter settings. The ratio of θ/ρ, which indicates the importance of mutation relative to recombination, was approximately 30–32, indicating that recombination, while present, typically causes substitutions much less frequently than mutation. It should be noted, however, that the mechanism of recombination among Ranavirus strains is not known, and so LDHat may not be able to accurately model the recombination process or estimate the rate of recombination. Number of recombination events detected by Rbrothers for several prior values Table 3 Number of recombination events detected by Rbrothers for several prior values Prior Expectation (“Lambda_Prior”) . Point Estimate . 95% Range . 1 50.5 39–56 5 65.8 60–72 10 67.1 61–74 20 79.4 71–88 Prior Expectation (“Lambda_Prior”) . Point Estimate . 95% Range . 1 50.5 39–56 5 65.8 60–72 10 67.1 61–74 20 79.4 71–88 Open in new tab Table 3 Number of recombination events detected by Rbrothers for several prior values Prior Expectation (“Lambda_Prior”) . Point Estimate . 95% Range . 1 50.5 39–56 5 65.8 60–72 10 67.1 61–74 20 79.4 71–88 Prior Expectation (“Lambda_Prior”) . Point Estimate . 95% Range . 1 50.5 39–56 5 65.8 60–72 10 67.1 61–74 20 79.4 71–88 Open in new tab Discussion We sequenced 15 strains of Ranavirus collected from Ambystoma tigrinum and A. mexicanum, collected from field sites and bait shops throughout western North America. Using comparative population genomic analyses, we characterized their diversity, identified potential targets of recent positive selection, and found evidence for recombination. Our results have the following implications: 1) the monophyly of ATV is upheld based on the genealogies constructed from complete genomes of all fully sequenced ranaviruses; 2) although there is evidence that ATV has been spread via recent movement of bait salamanders by humans, our best estimate of its original emergence time is several hundred years ago, before there were roads or a bait trade; 3) eight genes under strong positive selection were identified in ATV (however, the putative function of many of these genes is presently unknown); 4) ATV undergoes recombination, potentially allowing recombination between endemic and bait virus strains. Genomic diversity Although there is large-scale structural variation among Ranavirus genomes (Jancovich et al. 2010; Mavian et al. 2012a; see Figure 3), all of the ATV strains we sequenced were collinear with each other and with the reference ATV strain. Furthermore, there was little diversity in gene content among strains. This lack of diversity may be the result of similar selective forces on gene content—all of the strains sequenced here were collected from A. tigrinum in western North America—or the fairly recent origin of ATV, which may not have had time to accumulate gene content differences. Urodeles are known for relatively weak immune responses to viral infection (Cohen 1966; Chen et al. 2011); typical ATV die-offs occur in larval populations at ages when individuals cannot yet produce immune globulins (Brunner et al. 2004; Greer et al. 2008). In addition, a previous microarray study showed that ATV-infected larvae from A. mexicanum, the sister taxon of A. tigrinum, showed upregulation of genes associated with innate immunity, such as natural killer cells and cytokine signaling, but no appreciable change in expression of genes associated with adaptive immunity (Cotter et al. 2008). These implications are supported experimentally, as tiger salamanders are generally susceptible to ATV in infection trials (Storfer et al. 2007; Forson and Storfer 2006); susceptibility, however, is dose dependent (Brunner et al. 2005). As a result, there may not be strong selection by tiger salamander hosts for ATV to undergo radical shifts in genome structure or co-opt new genes. Although there was little diversity in gene content, there were three genes (ORF clusters 84, 85, and 88) present in all members of the ATV clade, but absent from all members other strains (see File S1 and Figure 2). These genes are potential adaptations to different host populations in different geographic areas, but none had a well-defined function. Cluster 84 was also identified as being under strong positive selection and showed homology with serum response factor-binding proteins, which are involved in regulation of the cell cycle (reviewed in Rodenberg et al. 2010). These three genes are attractive targets for future functional characterization, which may provide insight into the molecular mechanisms of host switches. Phylogeography Ranaviruses have been found in a wide range of cold-blooded vertebrates, and evidence suggests recent host switches among salamanders, turtles, frogs, and fish (Jancovich et al. 2010). Jancovich et al. (2010) found that ESV and EHNV, which were collected from fish, were more closely related to ATV than the frog or turtle ranaviruses. Our genealogy is consistent with their hypothesis that the common ancestor of the ranaviruses was likely a fish-infecting strain that experienced subsequent host shifts, although ambiguity in the placement of the root resulted in uncertainty about the placement of the ESV/EHNV clade relative to ATV. Jancovich et al. (2010) debated whether the common ancestor of the ALRV clade was an amphibian or a fish. Although our data cannot resolve this debate, there is evidence of additional host switching: ADRV, a salamander virus, is most closely related to CMTV, a toad virus (Balseiro et al. 2009; Mavian et al. 2012a). Note that the presence of recombination in this dataset means that no single genealogy can completely represent the evolutionary history of the samples. However, if recombination events are primarily old or take place between closely related individuals, the topology of the genealogy is expected to reflect the evolutionary history of the majority of the genome (Posada and Crandall 2002). We estimate that the recombination rate is low relative to mutation, and although we only tested for recombination within ATV, we expect that greater sequence divergence and lack of ecological opportunity (e.g., lack of geographic proximity or shared host) limit the amount of recombination between ATV and more distantly related samples. Thus, the conclusion that ATV is monophyletic should be robust to the presence of recombination, and the topology of the genealogies (Figure 2) likely represents the relationships across the majority of the genome. Our data also support the recent emergence of ATV: the upper limit of the 95% HPD interval for the age of the most recent common ancestor of the ATV strains was < 6000 years (point estimate = 545 years, 95% HPD range = 20–5672 years). The true emergence date likely lies between the most recent common ancestor and our estimated divergence date for ALRV (point estimate = 1526 years, 95% HPD = 53–16,524). Bayesian skyline analysis of all ALRVs (Figure 4) and just ATV (Figure S1) indicate no large changes in population sizes over the last several hundred years, although the confidence intervals are wide. These results suggest that, although recent geographic spread of ATV has resulted from human movement of infected salamanders in the bait trade (Jancovich et al. 2005; Picco and Collins 2008), the original emergence of ATV may have happened before such human influence, and any changes in effective population size that might be related to human activity have not yet produced a detectable signal in our data. Note that the emergence date we provide here is based on a small number of strains, all collected within the last 50 years, and so the confidence intervals are quite wide. In addition, recombination between strains can bias estimates of divergence dates: Schierup and Hein (2000) found a downward bias in divergence dates, though this was more pronounced for distance-based phylogeny methods than for maximum-likelihood methods. Thus, it is likely that our estimate of the emergence date of ATV is biased, probably downward. The Bayesian skyline analysis assumes no recombination (Drummond et al. 2005), and recombination tends to increase the lengths of the terminal branches of a tree, which can mimic the pattern produced by exponential population growth (Schierup and Hein 2000). Although our inferred rate of recombination is slow (about 1/30 of mutation), and we are not detecting evidence of exponential growth, it is possible that our estimated emergence dates are underestimates, and the effective population size estimates over time may be biased. Nonetheless, this result is consistent with comparative phylogenetic analysis of ATV and its tiger salamander hosts across western North America that showed strong evidence for genealogical concordance, and hence a historical coevolutionary relationship between this host and pathogen (Storfer et al. 2007). Selection To identify genes that may be particularly important for adaptation to new hosts or environments, we tested for positive selection using PAML and the MK test. After correcting for multiple comparisons, and removing nonfunctional sequences, eight genes showed evidence for positive selection within the ATV clade. Interestingly, two of these genes are potentially involved in immune response or viral replication: ORF cluster 84 has homology with serum response factor binding proteins, and ORF cluster 95 shows homology to collagen-like proteins, which may shield viral DNA from DNases in the host (Wang et al. 2013). ORF cluster 95 is also intriguing because it may be undergoing a process of pseudogenization in some strains, suggesting that it is not essential in all hosts or locations. In addition, the ORFs comprising cluster 95 have substantial genomic overlap with the ORFs comprising cluster 82. ORF cluster 93, a truncated 3-β-hydroxy-D 5-C27-steroid oxidoreductase-like protein (BOH) potentially involved in suppression of host immunity via upregulation of corticosteroids (Chinchar 2002; Delhon et al. 2006), is another gene showing evidence of pseudogenization, although after removal of the nonfunctional sequences, it does not appear to be subject to strong selection (note that the sequences included in cluster 93 were validated with Sanger sequencing data, and no errors were detected). In general, however, the genes with the strongest evidence for selection do not have well-characterized functions in the Ranavirus genus, so they are interesting targets for future work. It should be noted that the alignments of some of these genes contained substantial indel polymorphism (particularly ORF clusters 66, 84, and 95) that can cause misalignment and an inflated estimate of the number of nonsynonymous mutations. It is also notable that ORF clusters 9 (myristoylated membrane protein; orf53R in FV3), 62 (unknown; orf40R in FV3), and 75 (unknown; orf93L in FV3) were also identified as targets of strong positive selection by Abrams et al. (2013) with PAML analyses conducted across all of the ALRV. ORF cluster 9, a myristoylated membrane protein, is required for virion formation in FV3 (Whitley et al. 2010). That this gene appears to be a target of adaptive divergence between EHNV/ESV and ATV (indicated by the significant MK test) suggests possible importance for adaptation to new hosts or environments. We note that most ALRV genomes contain another copy of a myristoylated membrane protein (ORF cluster 55) that showed no evidence of positive selection. Similarly, ORF cluster 33 (18K immediate early protein) shows evidence of positive selection and is important for viral replication (Chen et al. 2011) (another immediate early protein—cluster 16—also shows evidence for positive selection), as is ORF cluster 1 (subunit a of DNA-dependent RNA polymerase). We also note that many genes may have positive selection on individual codons, but positive selection overall is not strong enough or widespread enough to reject the nearly neutral model across the entire gene. Recombination We provide one of the first tests for recombination among Ranavirus strains and the first study to quantify recombination using comparative genomic analyses. We found strong evidence for ongoing recombination within the ATV clade using several methods. Analyses using the RDP4 (Martin et al. 2010) and Rbrothers (Irvahn et al. 2013) packages, which include nine recombination-detection algorithms, all indicated the presence of recombination among ATV strains. Analyses using LDHat (Auton and McVean 2007) suggest the viral recombination rate, however, is slow (approximately 1/30 the rate of mutation) but sufficient to allow introgression between bait and wild strains. Although LDHat was designed for eukaryotes, and the physical mechanism of recombination in Ranavirus is not known, the estimates were very consistent across parameter values. For each putative recombination event, RDP4 attempts to identify a strain similar to the original nonrecombinant sequence (the “major parent”), and a strain similar to the donor of the recombinant sequence (the “minor parent”). It is worth noting that 2 of the 14 best-supported recombination events include potential movement of sequence from bait strains to wild strains (Table S1). However, identification of donor and recipient strains is challenging in a small, closely related sample, and manual examination of the RDP4 results indicates that the events likely represent ancestral recombination that occurred before human movement of salamanders or the evolution of bait strains. Nevertheless, these results have important implications for understanding the potential for pathogen pollution to affect ranavirus evolution in general and evolution of ATV in particular. Use of A. tigrinum as fishing bait has resulted in the movement of ATV strains over large distances, as well as the potential introduction of more virulent strains from bait farms or bait shops into wild populations (Jancovich et al. 2005; Storfer et al. 2007). Indeed, previous infection trials using bait vs. wild strains of ATV (Storfer et al. 2007) and FV3 (isolated from wild vs. farmed bullfrogs populations; A. Storfer and K. Chojnacki, unpublished data) have shown that bait strains cause greater mortality than native strains in infection trials. As a result, recombination among wild and bait-associated strains has the potential to introduce virulent alleles into locally adapted genomic backgrounds. These concerns extend to other ranavirus species, such as those in commercial aquaculture. Ranaviruses have caused die-offs in hatchery populations of rainbow trout (EHNV) as well as aquaculture populations of pallid sturgeon, European catfish (ECV) and grouper (GIV, SGIV) (Delhon et al. 2006), and when asymptomatic fish are used to stock natural populations, new viral strains could be introduced into naïve wild fish populations or recombine with existing strains. Acknowledgments We thank Rose Marie Larios and Karen Chojnacki for cultivating the viruses in cell culture and performing viral DNA extractions, and Jenn Cundiff for advice on viral DNA extractions. We also thank Matt Settles for assistance with genome assembly, Jesse Brunner for assistance with collecting strains, and Omar Cornejo for helpful comments that have improved the manuscript. This work was supported by National Science Foundation (NSF) grants DEB 0548415 and 1316549 from the NSF, and by the Washington State University Eastlick Distinguished Professorship to AS. Footnotes Supporting information is available online at www.g3journal.org/lookup/suppl/doi:10.1534/g3.115.023762/-/DC1 The data set generated for this article is available in GenBank: BioProject PRJNA257291. Communicating editor: Y. Kim Literature Cited Abrams A J , Cannatella D C, Hillis D M, Sawyer S L, 2013   Recent host-shifts in ranaviruses: signatures of positive selection in the viral genome. J. Gen. Virol. 94 : 2082 – 2093 . Google Scholar Crossref Search ADS PubMed WorldCat Angiuoli S V , Salzberg S L, 2011   Mugsy: fast multiple alignment of closely related whole genomes. Bioinformatics 27 : 334 – 342 . Google Scholar Crossref Search ADS PubMed WorldCat Auton A , McVean G, 2007   Recombination rate estimation in the presence of hotspots. Genome Res. 17 : 1219 – 1227 . Google Scholar Crossref Search ADS PubMed WorldCat Balseiro A , Dalton K P, del Cerro A, Marquez I, Cunningham A A et al. , 2009   Pathology, isolation and molecular characterisation of a ranavirus from the common midwife toad Alytes obstetricans on the Iberian Peninsula. Dis. Aquat. Organ. 84 : 95 – 104 . Google Scholar Crossref Search ADS PubMed WorldCat Bankevich A , Nurk S, Antipov D, Gurevich A A, Dvorkin M et al. , 2012   SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 19 : 455 – 477 . Google Scholar Crossref Search ADS PubMed WorldCat Bhatt S , Katzourakis A, Pybus O G, 2010   Detecting natural selection in RNA virus populations using sequence summary statistics. Infect. Genet. Evol. 10 : 421 – 430 . Google Scholar Crossref Search ADS PubMed WorldCat Boeckmann B , Bairoch A, Apweiler R, Blatter M C, Estreicher A et al. , 2003   The SWISS-PROT protein knowledgebase and its supplement TrEMBL in 2003. Nucleic Acids Res. 31 : 365 – 370 . Google Scholar Crossref Search ADS PubMed WorldCat Boni M F , Posada D, Feldman M W, 2007   An exact nonparametric method for inferring mosaic structure in sequence triplets. Genetics 176 : 1035 – 1047 . Google Scholar Crossref Search ADS PubMed WorldCat Brunner J , Schock D M, Davidson E W, Collins J P J P, 2004   Intraspecific reservoirs: complex life history and the persistence of a lethal ranavirus. Ecology 85 : 560 – 566 . Google Scholar Crossref Search ADS WorldCat Brunner J L , Richards K, Collins J P, 2005   Dose and host characteristics influence virulence of ranavirus infections. Oecologia 144 : 399 – 406 . Google Scholar Crossref Search ADS PubMed WorldCat Camacho C , Coulouris G, Avagyan V, Ma N, Papadopoulos J et al. , 2009   BLAST+: architecture and applications. BMC Bioinformatics 10 : 421 . Google Scholar Crossref Search ADS PubMed WorldCat Chen G , Ward B M, Yu K H, Chinchar V G, Robert J, 2011   Improved knockout methodology reveals that frog virus 3 mutants lacking either the 18K immediate-early gene of the truncated vIF-2α gene are defective for replication and growth in vivo. J. Virol. 85 : 11131 – 11138 . Google Scholar Crossref Search ADS PubMed WorldCat Chen Z-x , Zheng J-c, Jiang Y-l, 1999   A new iridovirus isolated from soft-shelled turtle. Virus Res. 63 : 147 – 151 . Google Scholar Crossref Search ADS PubMed WorldCat Chinchar V G , 2002   Ranaviruses (family Iridoviridae): emerging cold-blooded killers. Arch. Virol. 147 : 447 – 470 . Google Scholar Crossref Search ADS PubMed WorldCat Chinchar V G , Waltzek T B, 2014   Ranaviruses: not just for frogs. PLoS Pathog. 10 : e1003850 . Google Scholar Crossref Search ADS PubMed WorldCat Cohen N , 1966   Tissue transplantation immunity in the adult newt, Diemictylus viridescens. III. The effects of X-irradiation and temperature on the allograft reaction. J. Exp. Zool. 163 : 231 – 240 . Google Scholar Crossref Search ADS WorldCat Cotter J D , Storfer A, Page R B, Beachy C K, Voss S R, 2008   Transcriptional response of Mexican axlotls to Ambystoma tigrinum virus (ATV) infection. BMC Genomics 9 : 493 . Google Scholar Crossref Search ADS PubMed WorldCat Cunningham A A , Daszak P, Rodríguez J P, 2003   Pathogen pollution: defining a parasitological threat to biodiversity conservation. J. Parasitol. 89 : S78 – S83 . Google Scholar OpenURL Placeholder Text WorldCat Darriba D , Taboada G L, Doallo R, Posada D, 2012   jModelTest 2: more models, new heuristics and parallel computing. Nat. Methods 9 : 772 . Google Scholar Crossref Search ADS PubMed WorldCat Daszak P , Cunningham A A, Hyat A D, 2000   Emerging infectious diseases of wildlife—threats to biodiversity and human health. Science 287 : 443 – 449 . Google Scholar Crossref Search ADS PubMed WorldCat De Castro F , Bolker B, 2005   Mechanisms of disease-induced extinction. Ecol. Lett. 8 : 117 – 126 . Google Scholar Crossref Search ADS WorldCat Delhon G , Tulman E R, Afonso C L, Lu Z, Becnel J J et al. , 2006   Genome of invertebrate iridescent virus type 3 (mosquito iridescent virus). J. Virol. 80 : 8439 – 8449 . Google Scholar Crossref Search ADS PubMed WorldCat Drummond A J , Rambaut A, Shapiro B, Pybus O G, 2005   Bayesian coalescent inference of past population dynamics from molecular sequences. Mol. Biol. Evol. 22 : 1185 – 1192 . Google Scholar Crossref Search ADS PubMed WorldCat Drummond A J , Suchar M A, Xie D, Rambaut A, 2012   Bayesian phylogenetics with BEAUti and the BEAST 1. 7. Mol. Biol. Evol. 29 : 1969 – 1973 . Google Scholar Crossref Search ADS PubMed WorldCat Drummond A J , Ho S Y W, Phillips M J, Rambaut A, 2006   Relaxed phylogenetics and dating with confidence. PLoS Biol. 4 : e699 . Google Scholar Crossref Search ADS WorldCat Ebert D , 1999   The evolution and expression of parasite virulence , pp. 161 – 172 in Evolution in Health and Disease , edited by Stearns S C Oxford University Press , New York . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Edgar R C , 2004   MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32 : 1792 – 1797 . Google Scholar Crossref Search ADS PubMed WorldCat Fijan N , Sulimanovic D, Bearzotti M, Muzinic D, Zwillenberg L O et al. , 1983   Some properties of the Epithelioma papulosum cyprini (EPC) cell line from carp Cyprinus carpio. Ann. Inst. Pasteur Virol. 134 : 207 – 220 . Google Scholar Crossref Search ADS WorldCat Forson D D , Storfer A, 2006   Atrazine increases ranavirus susceptibility in the tiger salamander, Ambystoma tigrinum. Ecol. Appl. 16 : 2325 – 2332 . Google Scholar Crossref Search ADS PubMed WorldCat Geng Y , Wang K Y, Zhou Z Y, Li C W, Wang J et al. , 2011   First report of a Ranavirus associated with morbidity and mortality in farmed Chinese Giant Salamanders (Andrias davidianus). J. Comp. Pathol. 145 : 95 – 102 . Google Scholar Crossref Search ADS PubMed WorldCat Gibbs M J , Armstrong J S, Gibbs A J, 2000   Sister-scanning: a Monte Carlo procedure for assessing signals in recombinant sequences. Bioinformatics 16 : 573 – 582 . Google Scholar Crossref Search ADS PubMed WorldCat Granoff A , Came P E, 1965   The isolation and properties of viruses from Rana pipiens: their possible relationship to the renal adenocarcinoma of the leopard frog. Ann. N. Y. Acad. Sci. 126 : 237 – 255 . Google Scholar Crossref Search ADS PubMed WorldCat Greer A L , Briggs C J, Collins J P, 2008   Testing a key assumption of host-pathogen theory: density and disease transmission. Oikos 117 : 1667 – 1673 . Google Scholar Crossref Search ADS WorldCat He J G , Lü L M Deng H H, He S P, Weng et al., 2002   Sequence analysis of the complete genome of an iridovirus isolated from the tiger frog. Virology 292 : 185 – 197 . Google Scholar Crossref Search ADS PubMed WorldCat Huang Y , Huang X, Liu H, Gong J, Ouyang Z et al. , 2009   Complete sequence determination of a novel reptile iridovirus isolated from soft-shelled turtle and evolutionary analysis of Iridoviridae. BMC Genomics 10 : 224 . Google Scholar Crossref Search ADS PubMed WorldCat Hunter, S. S., R. T. Lyon, B. A. J. Sarver, K. Hardwick, L. J. Forney et al., 2015 Assembly by reduced complexity (ARC): a hybrid approach for targeted assembly of homologous sequences. bioRxiv. 014662. Available at: http://biorxiv.org/content/early/2015/02/07/014662.article-info. Accessed: 1 March 2015. Irvahn J , Chattopadhyay S, Sokurenko E V, Minin V N, 2013   rbrothers: R package for Bayesian multiple change-point recombination detection. Evol. Bioinform. 9 : 235 – 238 . Google Scholar Crossref Search ADS WorldCat Jancovich J K , Davidson E W, Morado J F, Jacobs B L, Collins J P, 1997   Isolation of a lethal virus from the endangered tiger salamander, Ambystoma tigrinum stebbinsi Lowe. Dis. Aquat. Organ. 31 : 161 – 167 . Google Scholar Crossref Search ADS WorldCat Jancovich J K , Davidson E W, Seiler A, Jacobs B L, Collins J P, 2001   Transmission of the Ambystoma tigrinum virus to alternative hosts. Dis. Aquat. Organ. 46 : 159 – 163 . Google Scholar Crossref Search ADS PubMed WorldCat Jancovich J K , Mao J, Chinchar V G, Wyatt C, Case S T et al. , 2003   Genomic sequence of a ranavirus (family Iridoviridae) associated with salamander mortalities in North America. Virology 316 : 90 – 103 . Google Scholar Crossref Search ADS PubMed WorldCat Jancovich J K , Davidson E W, Paramewaran N, Mao J, Chinchar V G et al. , 2005   Evidence for emergence of an amphibian iridoviral disease because of human-enhanced spread. Mol. Ecol. 14 : 213 – 224 . Google Scholar Crossref Search ADS PubMed WorldCat Jancovich J K , Bremont M, Touchman J W, Jacobs B L, 2010   Evidence for multiple recent host species shifts among the Ranaviruses (family Iridoviridae). Virology 84 : 2636 – 2647 . Google Scholar Crossref Search ADS WorldCat Jiang Y L , Zhang M, Gao L Y, 2011   Isolation and characterization of an iridovirus from sick giant salamander (Andrias davidianus). Chin. J. Virol. 27 : 274 – 282 . Google Scholar OpenURL Placeholder Text WorldCat Kimura M , 1968   Evolutionary rates at the molecular level. Nature 217 : 624 – 626 . Google Scholar Crossref Search ADS PubMed WorldCat Kurtz S , Phillippy A, Delcher A L, Smoot M, Shumway M et al. , 2004   Versatile and open software for comparing large genomes. Genome Biol. 5 : R12 . Google Scholar Crossref Search ADS PubMed WorldCat Lai Y-S , Murali S, Ju H-Y, Wu M-F, Guo I-C et al. , 2000   Two iridovirus-susceptible cell lines established from kidney and liver of grouper, Epinephelus awoara (Temminck & Schlegel), and partial characterisation of grouper iridovirus. J. Fish Dis. 23 : 379 – 388 . Google Scholar Crossref Search ADS WorldCat Langdon J S , Humphrey J D, Williams L M, 1986   First virus isolation from Australian fish: an irodovirus-like pathogen from redfin perch, Perca fluviatilis L. J. Fish Dis. 9 : 263 – 268 . Google Scholar Crossref Search ADS WorldCat Lei X-Y , Ou T, Zhu R-L, 2012   Sequencing and analysis of the complete genome of Rana grylio virus (RGV). Arch. Virol. 157 : 1559 – 1564 . Google Scholar Crossref Search ADS PubMed WorldCat Magoč T , Salzberg S L, 2011   FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 21 : 2957 – 2963 . Google Scholar OpenURL Placeholder Text WorldCat Marguiles M , Egholm M, Altman W E, Attiya S, Bader J S et al. , 2005   Genome sequencing in microfabricated high-density picolitre reactors. Nature 437 : 376 – 381 . Google Scholar Crossref Search ADS PubMed WorldCat Martin D , Rybicki E, 2000   RDP: detection of recombination amongst aligned sequences. Bioinformatics 16 : 562 – 563 . Google Scholar Crossref Search ADS PubMed WorldCat Martin D P , Posada D, Crandall K A, Williamson C, 2005   A modified bootscan algorithm for automated identification of recombinant sequences and recombination breakpoints. AIDS Res. Hum. Retroviruses 21 : 98 – 102 . Google Scholar Crossref Search ADS PubMed WorldCat Martin D P , Lemey P, Lott M, Moulton V, Posada D et al. , 2010   RDP3: a flexible and fast computer program for analyzing recombination. Bioinformatics 26 : 2462 – 2463 . Google Scholar Crossref Search ADS PubMed WorldCat Mavian C , López-Bueno A, Balseiro A, Casais R, Alcamí A et al. , 2012 a   The genome sequence of the emerging common midwife toad virus identifies an evolutionary intermediate within Ranaviruses. J. Virol. 86 : 3617 – 3625 . Google Scholar Crossref Search ADS PubMed WorldCat Mavian C , López-Bueno A, Somalo M P F, Alcamí A, Alejo A, 2012 b   Complete genome sequence of the European Sheatfish Virus. J. Virol. 86 : 6365 . Google Scholar Crossref Search ADS PubMed WorldCat McDonald J H , Kreitman M, 1991   Adaptive protein evolution at the Adh locus in Drosophila. Nature 351 : 652 – 654 . Google Scholar Crossref Search ADS PubMed WorldCat Miller D , Gray M, Storfer A, 2011   Ecopathology of Ranaviruses infecting amphibians. Viruses 3 : 2351 – 2373 . Google Scholar Crossref Search ADS PubMed WorldCat Minin V N , Dorman K S, Fang F, Suchard M A, 2005   Dual multiple change-point model leads to more accurate recombination detection. Bioinformatics 21 : 3034 – 3042 . Google Scholar Crossref Search ADS PubMed WorldCat Otto T D , Dillon G P, Degrave W S, Berriman M, 2011   RATT . Rapid Annotation Transfer Tool. Nucleic Acids Res. 39 : e57 . Google Scholar Crossref Search ADS WorldCat Padidam M , Sawyer S, Fauquet C M, 1999   Possible emergence of new geminiviruses by frequent recombination. Virology 265 : 218 – 225 . Google Scholar Crossref Search ADS PubMed WorldCat Picco A M , Collins J P, 2008   Amphibian commerce as a likely source of pathogen pollution. Conserv. Biol. 22 : 1582 – 1589 . Google Scholar Crossref Search ADS PubMed WorldCat Posada D , Crandall K A, 2001   Evaluation of methods for detecting recombination from DNA sequences: computer simulations. Proc. Natl. Acad. USA. 98 : 13757 – 13762 . Google Scholar Crossref Search ADS WorldCat Posada D , Crandall K A, 2002   The effect of recombination on the accuracy of phylogeny estimation. J. Mol. Evol. 54 : 396 – 402 . Google Scholar Crossref Search ADS PubMed WorldCat Qin Q W , Lam T J, Sin Y M, Shen H, Chang S F et al. , 2001   Electron microscopic observations of a marine fish iridovirus isolated from brown-spotted grouper, Epinephelus tauvina. J. Virol. Methods 98 : 17 – 24 . Google Scholar Crossref Search ADS PubMed WorldCat R Core Team , 2014   R: language and environment for statistical computing , R Foundation for Statistical Computing , Vienna, Austria . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Ridenhour B J , Storfer A T, 2008   Geographically variable selection in Ambystoma tigrinum virus (Iridoviridae) throughout the western USA. J. Evol. Biol. 21 : 1151 – 1159 . Google Scholar Crossref Search ADS PubMed WorldCat Rodenberg J M , Hoggat A M, Chen M, Touw K, Jones R et al. , 2010   Regulation of serum response factor activity and smooth muscle cell apoptosis by chromodomain helicase DNA-binding protein 8. Am. J. Physiol. Cell Physiol. 299 : C1058 – C1067 . Google Scholar Crossref Search ADS PubMed WorldCat Ronquist F , Teslenko M, van der Mark P, Ayres D L, Darling A et al. , 2012   MrBayes 3. 2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61 : 539 – 542 . Google Scholar Crossref Search ADS PubMed WorldCat Schierup M H , Hein J, 2000   Consequences of recombination on traditional phylogenetic analysis. Genetics 156 : 879 – 891 . Google Scholar PubMed OpenURL Placeholder Text WorldCat Schloegel L M , Picco A M, Kilpatrick A M, Davies A J, Hyatt A D et al. , 2009   Magnitude of the US trade in amphibians and presence of Batrachochytrium dendrobatidis and ranavirus infection in imported North American bullfrogs (Rana catesbiana). Biol. Conserv. 142 : 1420 . Google Scholar Crossref Search ADS WorldCat Schloegel L M , Daszak P, Cunningham A A, Speare R, Hill B, 2010   Two amphibian diseases, chytridiomycosis and ranaviral disease, are now globally notifiable to the world organization for animal health (OIE): an assessment. Dis. Aquat. Organ. 92 : 101 – 108 . Google Scholar Crossref Search ADS PubMed WorldCat Smith G J D , Vijaykrishna D, Bahl J, Lycett S J, Worobey M et al. , 2009   Origins and evolutionary genomics of the 2009 swine-origin H1N1 influenza A epidemic. Nature 459 : 1122 – 1126 . Google Scholar Crossref Search ADS PubMed WorldCat Smith J M , 1992   Analyzing the mosaic structure of genes. J. Mol. Evol. 34 : 126 – 129 . Google Scholar PubMed OpenURL Placeholder Text WorldCat Smith K F , Sax D F, Lafferty K D, 2006   Evidence for the role of infectious disease in species extinction and endangerment. Conserv. Biol. 20 : 1349 – 1357 . Google Scholar Crossref Search ADS PubMed WorldCat Song W J , Qin Q W, Qui J, Huang C H, Wang F et al. , 2004   Functional genomics analysis of Singapore Grouper Iridovirus: complete sequence determination and proteomic analysis. J. Virol. 78 : 12576 – 12590 . Google Scholar Crossref Search ADS PubMed WorldCat Stöhr A C , López-Bueno A, Blahak S, Caeiro M F, Rosa G M et al. , 2015   Phylogeny and differentiation of reptilian and amphibian ranaviruses detected in Europe. PLoS One 10 : e0118633 . Google Scholar Crossref Search ADS PubMed WorldCat Storfer A , Alfaro M E, Ridenhour B J, Jancovich J K, Mech S G et al. , 2007   Phylogenetic concordance analysis shows an emerging pathogen is novel and endemic. Ecol. Lett. 10 : 1075 – 1083 . Google Scholar Crossref Search ADS PubMed WorldCat Tajima F , 1989   Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123 : 585 – 595 . Google Scholar Crossref Search ADS PubMed WorldCat Tan W G H , Barkman T J, Chinchar V G, Essani K, 2004   Comparative genomic analyses of frog virus 3, type species of the genus Ranavirus (family Iridoviridae). J. Virol. 323 : 70 – 84 . Google Scholar Crossref Search ADS WorldCat Thornton K , 2003   libsequence: a C++ class library for evolutionary genetic analysis. Bioinformatics 19 : 2325 – 2327 . Google Scholar Crossref Search ADS PubMed WorldCat Thurber R V , Haynes M, Breitbart M, Wegley L, Rohwer F, 2009   Laboratory procedures to generate viral metagenomes. Nat. Protoc. 4 : 470 – 483 . Google Scholar Crossref Search ADS PubMed WorldCat Tsai C-T , Ting J-W, Wu M-H, Wu M-F, Guo I-C et al. , 2005   Complete genome sequence of the grouper iridovirus and comparison of genomic organization with those of other iridoviruses. J. Virol. 79 : 2010 – 2023 . Google Scholar Crossref Search ADS PubMed WorldCat Wang L , Chong Q Y, Wu J, 2013   A DNA-binding protein encoded by ORF008L of the Singapore grouper iridovirus. Virus Res. 176 : 37 – 44 . Google Scholar Crossref Search ADS PubMed WorldCat Wang N , Zhang M, Zhang L, Jing H, Jiang Y et al. , 2014   Complete genome sequence of a Ranavirus isolated from chinese giant salamander (Andrias davidianus). Genome Announc. 2 : e01032 – e13 . Google Scholar PubMed OpenURL Placeholder Text WorldCat Watterson G A , 1975   On the number of segregating sites in genetical models without recombination. Theor. Popul. Biol. 7 : 256 – 276 . Google Scholar Crossref Search ADS PubMed WorldCat Whitley D S , Yu K, Sample R C, Sinning A, Henegar J et al. , 2010   Frog virus 3 ORF 53R, a putative myristoylated membrane protein, is essential for virus replication in vitro. Virology 405 : 448 – 456 . Google Scholar Crossref Search ADS PubMed WorldCat Yang Z , 2007   PAML 4: Phylogenetic analysis by maximum likelihood. Mol. Biol. Evol. 24 : 1586 – 1591 . Google Scholar Crossref Search ADS PubMed WorldCat Zhang Q-Y , Xiao F, Li Z-Q, Gui J-F, Mao J et al. , 2001   Characterization of an iridovirus from the cultured pig frog Rana grylio with lethal syndrome. Dis. Aquat. Organ. 48 : 27 – 36 . Google Scholar Crossref Search ADS PubMed WorldCat © 2016 Epstein, Storfer This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model) © 2016 Epstein, Storfer TI - Comparative Genomics of an Emerging Amphibian Virus JF - "G3: Genes, Genomes, Genetics" DO - 10.1534/g3.115.023762 DA - 2016-01-01 UR - https://www.deepdyve.com/lp/oxford-university-press/comparative-genomics-of-an-emerging-amphibian-virus-BL9viZZ0F0 SP - 15 EP - 27 VL - 6 IS - 1 DP - DeepDyve ER -