Abstract

Antarctica harbours some of the most isolated and extreme environments on Earth, concealing a largely unexplored and unique component of the global animal virosphere. To understand the diversity and evolutionary histories of viruses in these polar species, we determined the viromes of gill metatranscriptomes from 11 Antarctic fish species with 248 samples collected from the Ross Sea region spanning the Perciformes, Gadiformes, and Scorpaeniformes orders. The continent’s shift southward and cooling temperatures >20 million years ago led to a reduction in biodiversity and subsequent radiation of some marine fauna, such as the notothenioid fishes. Despite decreased host species richness in polar regions, we revealed a surprisingly complex virome diversity in Ross Sea fish, with the types and numbers of viruses per host species and individuals sampled comparable to that of fish in warmer marine environments with higher host community diversity. We also observed a higher number of closely related viruses likely representing instances of recent and historic host-switching events among the Perciformes (all notothenioids) than in the Gadiformes, suggesting that rapid speciation events within this order generated closely related host species with few genetic barriers to cross-species transmission. Additionally, we identified novel genomic variation in an arenavirus with a split nucleoprotein sequence containing a stable helical structure, indicating potential adaptation of viral proteins to extreme temperatures. These findings enhance our understanding of virus evolution and virus-host interactions in response to environmental shifts, especially in less diverse ecosystems that are more vulnerable to the impacts of anthropogenic and climate changes.

Introduction

Antarctica’s icy waters hide some of the world’s most isolated and hostile environments, and with it, its aquatic life shields a unique and largely unexplored pocket of the vertebrate virosphere. As the continent shifted southward >20 million years ago, the cooling of the Southern Ocean towards an increasingly polar climate saw a reduction in the temperate Eocene fauna that previously inhabited Antarctica and opened increasingly extreme and isolated niches, reshaping the diversity of its marine life (Eastman 2005, Daane and Detrich 2022). Water temperatures in the Southern Ocean consistently sit below the freezing point of fish blood (Littlepage 2013). Ray-finned fishes from the Notothenioidei, Liparidae, and Zoarcoidei account for most of the fish diversity in Antarctica (Eastman 2005), and unique adaptations—such as antifreeze glycoproteins (AFGPs) in their blood—allowed them to prevail in these subzero climates where other temperate species could not (Devries 1971). Subsequent specialization to the changing environment, including instances of radiation, particularly of the notothenioid species (order: Perciformes), has created a taxonomically restricted modern ichthyofauna (Clarke and Johnston 1996, Eastman and Clarke 1998). Today, there are ∼320 known species of Antarctic fish (Eastman 2005), a relatively small number given that fish compose >50% of Earth’s vertebrate species (Nelson 2006). Notothenioids alone have radiated to >130 species and make up >90% of the fish biomass in the Southern Ocean (Eastman and Hubold 1999, Daane and Detrich 2022). Radiation of Antarctic fishes has also resulted in the highest rates of endemism of any isolated marine environment (Eastman and Clarke 1998). Such species not only make for fascinating models of evolution (Seehausen 2006, Thacker et al. 2021) but also for studying their associated microbial flora, particularly viruses (Perry et al. 2022, Costa et al. 2023b). Still, much of the focus on the viruses of Antarctic species to date has revolved around birds, seals (Varsani et al. 2017, Smeele et al. 2018), environmental samples (Aguirre de Cárcer et al. 2016, Gong et al. 2018), as well as Trematomus fish (Kraberger et al. 2022).

Growing attention to these polar species has also left them vulnerable to several threats. Antarctic fish, and their AFGPs, are scientifically, economically, and commercially important (Collins et al. 2010, Eskandari et al. 2020). However, overfishing in the last 50 years has induced concerns over population recovery, bycatch, and competition from marine predators. This is particularly acute for highly sought-after species like the Patagonian toothfish (Dissostichus eleginoides) and common bycatch such as grenadier and Antimora rostrata (Collins et al. 2010, Smith et al. 2012) for which little is known about their biology and ecology. Antarctic fish are also sensitive to environmental change. The Southern Ocean has been warming considerably faster than the global ocean (Gille 2002) and unlike Antarctica’s speciose tropical counterparts, the lack of biodiversity in this region leaves it without a protective buffer making it less resilient to climatic and anthropogenic changes (Duffy et al. 2016). Mortality of fish due to viral infections and modulation of host immune responses have also been shown to be affected by increasing temperatures (Jung et al. 2017, Páez et al. 2021). Hence, a better understanding of the biology, ecology, and microbes of polar fish may be important in the management and protection of these species in the face of these growing issues.

Species richness increases towards the equator (Rabosky et al. 2018) and this diversity has recently been shown to be equally manifested in that of fish viromes in these warmer waters (Costa et al. 2023a). Due to the relatively restricted and specialist aquatic taxonomy in the Southern Ocean, it may be expected that the viromes of these species are similarly restricted or divergent in comparison to those in more equatorial ecosystems. The Ross Sea is the largest embayment in Antarctica at 180° S and contains the world’s largest marine protected area (Smith et al. 2012, Cummings et al. 2021). At least 80 fish species reside in the Ross Sea, dominated by the notothenioids (making up >60% of species) and a number of non-notothenioid species (Eastman and Hubold 1999). To better understand the viromes of these lesser-studied aquatic species, we sampled Antarctic fishes from the Ross Sea, spanning the Perciformes, Gadiformes, and Scorpaeniformes orders. Our aim was to use metatranscriptomic sequencing to examine their virome diversity and determine how historic oceanographic, evolutionary, and climatic changes may have shaped the diversity and evolution of these viruses compared to those of more temperate or tropical species.

Materials and methods

Ethics

Samples were collected under the San Aotea II 2022 AMLR permit, under the Ross Sea data collection plan, the Ross Sea shelf survey plan, and in accordance with the relevant Commission for the Conservation of Antarctic Marine Living Resources (CCAMLR) Conservation Measures.

Gill sample collection

Antarctic fish were collected between early December 2021 and mid-January 2022 by the San Aotea II during commercial fishing in the Ross Sea and research fishing on the Ross Sea shelf survey at depths of 522–1810 m. Species were identified where possible or placed into the next most descriptive taxonomic group (family or genus). Importantly, all of the fish were in good condition with no obvious signs of disease. Whole fish heads were frozen and stored at −4°C ready for dissection of gill tissue.

Frozen heads were partially defrosted and ∼0.5–1 cm3 of gill tissue was dissected using sterilized scalpels and forceps and stored in 1 ml of RNAlater (ThermoFisher Scientific) before storage at −80°C. Gill samples were taken from between 4 and 23 individuals per species group for a total of 248 samples from three orders: Perciformes (n = 118) (Chionobathyscus dewitti, Trematomus loennbergii, Trematomus lepidorhinus, Pogonophryne barsukovi, Pogonophryne immaculata, and Pogonophryne scotti), Gadiformes (n = 121) (Macrourus caml, Macrourus whitsoni, A. rostrata, and Muraenolepis spp.), and Scorpaeniformes (n = 9) (Zoarcidae spp.) (Table 1).

Table 1.

Overview of sampling libraries.

LibraryHost species name/groupTaxonomic orderNumber of fishBody mass (kg)
MRL_LMuraenolepis spp.Gadiformes110.51–0.74
MRL_HMuraenolepis spp.Gadiformes80.77–1.01
ANT_LAntimora rostrataGadiformes120.98–1.91
ANT_HAntimora rostrataGadiformes112.02–3.72
WGR_LMacrourus whitsoniGadiformes180.21–0.58
WGR_HMacrourus whitsoniGadiformes200.6–1.28
QMC_LMacrourus camlGadiformes230.67–1.26
QMC_HMacrourus camlGadiformes181.24–2.25
LVD_LZoarcidaeScorpaeniformes50.04–0.1
LVD_HZoarcidaeScorpaeniformes40.11–0.22
TRD_LTrematomus lepidorhinusPerciformes110.14–0.32
TRD_HTrematomus lepidorhinusPerciformes80.33–0.5
TLO_LTrematomus loennbergiiPerciformes100.19–0.35
TLO_HTrematomus loennbergiiPerciformes70.36–0.52
CHW_LChionobathyscus dewittiPerciformes210.17–0.29
CHW_HChionobathyscus dewittiPerciformes190.3–0.68
SZT_LPogonophryne scottiPerciformes50.19–0.26
SZT_LPogonophryne scottiPerciformes50.3–0.43
PIM_LPogonophryne immaculataPerciformes90.15–0.24
PIM_HPogonophryne immaculataPerciformes90.26–0.36
BVK_LPogonophryne barsukoviPerciformes70.26–0.38
BVK_HPogonophryne barsukoviPerciformes70.4–0.5
LibraryHost species name/groupTaxonomic orderNumber of fishBody mass (kg)
MRL_LMuraenolepis spp.Gadiformes110.51–0.74
MRL_HMuraenolepis spp.Gadiformes80.77–1.01
ANT_LAntimora rostrataGadiformes120.98–1.91
ANT_HAntimora rostrataGadiformes112.02–3.72
WGR_LMacrourus whitsoniGadiformes180.21–0.58
WGR_HMacrourus whitsoniGadiformes200.6–1.28
QMC_LMacrourus camlGadiformes230.67–1.26
QMC_HMacrourus camlGadiformes181.24–2.25
LVD_LZoarcidaeScorpaeniformes50.04–0.1
LVD_HZoarcidaeScorpaeniformes40.11–0.22
TRD_LTrematomus lepidorhinusPerciformes110.14–0.32
TRD_HTrematomus lepidorhinusPerciformes80.33–0.5
TLO_LTrematomus loennbergiiPerciformes100.19–0.35
TLO_HTrematomus loennbergiiPerciformes70.36–0.52
CHW_LChionobathyscus dewittiPerciformes210.17–0.29
CHW_HChionobathyscus dewittiPerciformes190.3–0.68
SZT_LPogonophryne scottiPerciformes50.19–0.26
SZT_LPogonophryne scottiPerciformes50.3–0.43
PIM_LPogonophryne immaculataPerciformes90.15–0.24
PIM_HPogonophryne immaculataPerciformes90.26–0.36
BVK_LPogonophryne barsukoviPerciformes70.26–0.38
BVK_HPogonophryne barsukoviPerciformes70.4–0.5
Table 1.

Overview of sampling libraries.

LibraryHost species name/groupTaxonomic orderNumber of fishBody mass (kg)
MRL_LMuraenolepis spp.Gadiformes110.51–0.74
MRL_HMuraenolepis spp.Gadiformes80.77–1.01
ANT_LAntimora rostrataGadiformes120.98–1.91
ANT_HAntimora rostrataGadiformes112.02–3.72
WGR_LMacrourus whitsoniGadiformes180.21–0.58
WGR_HMacrourus whitsoniGadiformes200.6–1.28
QMC_LMacrourus camlGadiformes230.67–1.26
QMC_HMacrourus camlGadiformes181.24–2.25
LVD_LZoarcidaeScorpaeniformes50.04–0.1
LVD_HZoarcidaeScorpaeniformes40.11–0.22
TRD_LTrematomus lepidorhinusPerciformes110.14–0.32
TRD_HTrematomus lepidorhinusPerciformes80.33–0.5
TLO_LTrematomus loennbergiiPerciformes100.19–0.35
TLO_HTrematomus loennbergiiPerciformes70.36–0.52
CHW_LChionobathyscus dewittiPerciformes210.17–0.29
CHW_HChionobathyscus dewittiPerciformes190.3–0.68
SZT_LPogonophryne scottiPerciformes50.19–0.26
SZT_LPogonophryne scottiPerciformes50.3–0.43
PIM_LPogonophryne immaculataPerciformes90.15–0.24
PIM_HPogonophryne immaculataPerciformes90.26–0.36
BVK_LPogonophryne barsukoviPerciformes70.26–0.38
BVK_HPogonophryne barsukoviPerciformes70.4–0.5
LibraryHost species name/groupTaxonomic orderNumber of fishBody mass (kg)
MRL_LMuraenolepis spp.Gadiformes110.51–0.74
MRL_HMuraenolepis spp.Gadiformes80.77–1.01
ANT_LAntimora rostrataGadiformes120.98–1.91
ANT_HAntimora rostrataGadiformes112.02–3.72
WGR_LMacrourus whitsoniGadiformes180.21–0.58
WGR_HMacrourus whitsoniGadiformes200.6–1.28
QMC_LMacrourus camlGadiformes230.67–1.26
QMC_HMacrourus camlGadiformes181.24–2.25
LVD_LZoarcidaeScorpaeniformes50.04–0.1
LVD_HZoarcidaeScorpaeniformes40.11–0.22
TRD_LTrematomus lepidorhinusPerciformes110.14–0.32
TRD_HTrematomus lepidorhinusPerciformes80.33–0.5
TLO_LTrematomus loennbergiiPerciformes100.19–0.35
TLO_HTrematomus loennbergiiPerciformes70.36–0.52
CHW_LChionobathyscus dewittiPerciformes210.17–0.29
CHW_HChionobathyscus dewittiPerciformes190.3–0.68
SZT_LPogonophryne scottiPerciformes50.19–0.26
SZT_LPogonophryne scottiPerciformes50.3–0.43
PIM_LPogonophryne immaculataPerciformes90.15–0.24
PIM_HPogonophryne immaculataPerciformes90.26–0.36
BVK_LPogonophryne barsukoviPerciformes70.26–0.38
BVK_HPogonophryne barsukoviPerciformes70.4–0.5

RNA extraction and sequencing

For RNA extractions, the RNeasy Plus Mini Kit (Qiagen) protocol was followed for all samples. Frozen gill tissue was initially prepared by partially thawing and ∼10 mg was taken using sterilized scalpels and forceps and placed into Buffer RLT lysis buffer containing 1% β-mercaptoethanol and 0.5% (v/v) Reagent DX. A Qiagen TissueRuptor was used to homogenize the tissue in the lysis buffer for ∼30 s per sample. The resulting homogenates were centrifuged to remove residual tissue and the protocol was resumed to extract whole RNA. RNA concentrations were quantified using a NanoDrop Spectrophotometer (ThermoFisher).

RNA (5–20 μl) from each individual was pooled by species and into either ‘low’ (L) or ‘high’ (H) body mass (kg) for a total of 22 library pools containing RNA from 4 to 23 fish each (Table 1).

Illumina Stranded Total RNA Prep with Ribo-Zero plus (Illumina) was used for library preparation of the 22 pooled RNA libraries and these were subjected to total RNA paired-end sequencing on the Illumina NovaSeq 6000 platform, generating 150-bp reads.

Fish gill metatranscriptome assembly and annotation

Trinity RNA-seq (v2.11) (Grabherr et al. 2011) was used for de novo assembly of the gill metatranscriptomes. Default parameters for paired-end read inputs were used with the additional ‘trimmomatic’ flag option to perform pre-assembly quality trimming of the reads. Trinity contigs from each library were annotated both at the nucleotide level against the National Center for Biotechnology Information (NCBI)’s nucleotide (nt) database using the BLASTn algorithm (Camacho et al. 2009) and at the amino acid level against the non-redundant (nr) protein database using DIAMOND BLASTx (v2.02.2) (Buchfink et al. 2021).

Vertebrate virus discovery and abundance estimation

Contigs annotated as being potentially viral in origin in one or both of the nucleotide or protein search outputs with e-values of <1 × 10−10 were checked with additional blastn and blastp searches using the online BLAST server (https://blast.ncbi.nlm.nih.gov/Blast.cgi) to eliminate false positive hits. Sequences with e-values of >1 × 10−10 or where the top blastn or blastp hits included non-viral or non-vertebrate viral species based on their taxonomic assignments and/or proposed hosts (e.g. bacteriophages, environmental metagenome-, fungi-, plant-, or invertebrate-associated viruses) were excluded from further analyses. Putative vertebrate virus sequences were considered if their top blastp hits were to those of other previously known vertebrate viruses. DNA virus sequences were also checked with additional blastn searches to eliminate those sequences with homology to fish genomic DNA to exclude possible endogenous viral elements.

Abundance estimations of contigs were generated using the ‘align and estimate’ module within Trinity with the ‘prep reference’ flag set, RNA-seq by Expectation-Maximization (RSEM) (Li and Dewey 2011) as the abundance estimation method, and Bowtie 2 (Langmead and Salzberg 2012) as the alignment method. Contig abundances were standardized for inter-library comparisons by dividing RSEM abundance counts by the total sequencing read depth in their respective libraries. Standardized abundances were converted to reads-per-million (RPM) by multiplying by 1 million. To reduce incorrect assignment of viruses to libraries due to index-hopping, virus sequences sharing >99% nucleotide identity with sequences in other libraries occurring at abundances <0.1% of the highest abundance for that virus across all libraries were considered contamination and were removed from subsequent analyses.

For comparison of the presence and abundance of viruses between host taxonomic groups, abundances from each of the two body mass libraries per host group were merged. Abundances of individual viruses were compared alongside the standardized abundance of the stably expressed host gene, 40S Ribosomal Protein S13 (RPS13). RPS13 shows minimal inter- and intra-tissue expression-level variation, making it a suitable reference host gene for transcript abundance comparisons (Robichaux et al. 2016). Where full genomes or genome segments of novel viruses were suspected due to the presence of full open reading frames (ORFs) of the expected lengths for viruses in their respective families, all novel virus segments for that virus were aligned with those from their closest known relatives for confirmation.

Phylogenetic analysis

Inferences about the evolutionary relationships of the viruses documented here were made using maximum likelihood trees generated with IQ-TREE (v1.6.12) (Nguyen et al. 2015). We assumed that the viruses identified here that clustered with known viruses from other fish or vertebrate host species in their respective phylogenies were likely to be replicating in the fish host directly, rather than being associated with fish diets, microbiomes, or environments. All phylogenies were generated using amino acid sequences containing the highly conserved polymerase genes [either the RNA-dependent RNA polymerase (RdRp) for RNA viruses or the DNA-dependent DNA polymerase (DdDp) for DNA viruses]. Viruses that could be assigned down to at least order level based on shared sequence identity with their closest relatives identified by BLAST searches were collated with a range of viruses from their respective taxonomies retrieved from NCBI Taxonomy (https://www.ncbi.nlm.nih.gov/Taxonomy/Browser/wwwtax.cgi) and aligned using the MAFFT (v7.450) (Katoh and Standley 2013) L-INS-i algorithm. The exceptions were viruses from the Nackednaviridae and Hepadnaviridae families, which were aligned together with sequences from Lauber et al. (2017) rather than from NCBI Taxonomy. Alignments were manually inspected in Geneious Prime (v2023.2.1) (https://www.geneious.com/) to observe the expected alignment of conserved amino acid motifs and ambiguously aligned regions were trimmed using trimAl (V1.2) (Capella-Gutiérrez et al. 2009) with the ‘automated1’ flag set, or in the case of the Picornaviridae, with the flags ‘gt’ (gap threshold) set to 0.9 and ‘cons’ (variable conserve value) set to 50 in order to prevent excessive trimming. Sequence alignments for each taxonomic group were used to estimate phylogenies using IQ-TREE with the LG amino acid substitution model and 1000 ultra-fast bootstrapping replicates (Hoang et al. 2018). The ‘alrt’ flag was also added to perform 1000 bootstrap replicates for the Shimodaira-Hasegawa-like approximate likelihood ratio test (Guindon et al. 2010). Phylogenetic trees were annotated in FigTree (v1.4.4) (http://tree.bio.ed.ac.uk/software/figtree/) and rooted at their midpoints.

For viruses of the Arenaviridae found in pools of Trematomous fish in this study and through data mining (as mentioned later), phylogenetic trees were also produced using amino acid sequences of nucleoproteins (NPs) using the same method above.

Data mining for arenaviruses and virus genome segment recovery

To identify further arenaviruses that may have unique genomic and structural variations present in the NCBI Transcriptome Shotgun Assembly (TSA) database, full-length amino acid sequences of the L protein (polymerase), segmented NP, and glycoprotein from Trematomous arenavirus (PP590693 and PP590768-9) were used as bait to screen all vertebrate transcriptome assemblies (taxonomic identifier: 7742), excluding Homo sapiens (taxonomic identifier: 9606), using the online translated Basic Local Alignment nucleotide (tblastn) search tool. The BLOSUM45 matrix was used to increase the chance of finding highly divergent viruses. Any putative viral hits to the ‘bait’ sequences were queried using further blastp searches for confirmation. Raw read data associated with all positive hits were recovered from the Short Read Archive (SRA) as follows: SRR3184758 (resting phase testis of Channa punctata), SRR12526228 (RNAseq of Coregonus artedi), and SRR2912518 (transcriptome sequencing of white sucker (Catostomus commersonii)). Trinity RNA-seq (see Fish gill metatranscriptome assembly and annotation) for SRR318478 and MEGAHIT (v1.2.9) (Li et al. 2015) for SRR12526228 and SRR2912518 with default parameters were used to reassemble transcriptomes of the mined host data in an attempt to recover all possible genome segments for each new arenavirus identified. Assembled transcriptomes were screened against the nr protein database using DIAMOND BLASTx searches and the outputs were manually screened for the sequences of interest.

NP structural analysis

To explore the segmented NP of the novel Trematomous arenavirus, structures of the full coding regions (ORFs) of NP segments were predicted using the Google Colab implementation of AlphaFold2 using MMseq2 with default settings (Mirdita et al. 2022). The resulting structural models were compared against the Protein Data Bank using the PDB search on the online Dali server (Holm 2022) to identify which segments of arenavirus NPs the novel structures represented. Relevant structural superimpositions of novel NPs with those of a representative arenavirus NP structure, Lassa virus (Protein Data Bank: 3MX5), and all figures of structures were prepared using UCSF ChimeraX (Pettersen et al. 2021).

Testing effects of host factors on virome composition

The potential effects of host taxonomy and body mass group on virome composition were tested using various packages in R (v4.1.1). Standardized virome abundances were normalized and a distance matrix was created using the vdist function from the vegan package (Oksanen et al. 2022). Bray–Curtis dissimilarity as the distance measure and non-metric multidimensional scaling was performed on the distance matrix using the metaMDS function in vegan. Permutational multivariate analysis of variance was used to test for statistical significance of the effect of host taxonomy (order and genus) and body mass group (high or low, see Table 1) on virome composition (presence and abundance) using the adonis2 function in vegan. Data were plotted using ggplot2 (Wickham 2016).

Virus nomenclature

Viruses were considered novel if they shared <90% RdRp or DdDp amino acid identity with their closest known relative or <80% genome (nucleotide) identity with previously described virus species. This is a general classification based on the International Committee on Taxonomy of Viruses guides for the classification of novel viruses for some taxonomic groups (Lefkowitz et al. 2018), as many have no or unclear genomic thresholds for proposing novelty, and the presumption that viruses with <90% amino acid identity in their most conserved proteins are unlikely to represent previously known species. Putative viruses were provisionally named according to either the common or scientific name of their host group.

Data availability

Raw sequence reads are available on the NCBI SRA under the BioProject accession PRJNA1088854. Virus genome sequences identified in this study are available under the GenBank accessions PP590684–PP590779 (Supplementary Table S2). Extended data outlining additional (non-polymerase) virus segments, lengths, and sequences from all of the viruses identified in this study, as well as R code and accompanying input data used to generate results and figures in this study, can be found on GitHub: https://github.com/maybec49/Antarctic_Fish_Viromes.

Results

Virome diversity and abundance in Antarctic fish

Gill metatranscriptomes yielded ∼61–119 million sequencing reads per library (Fig. 1) and de novo assembly generated 473 258–782 113 contigs per library. Total virome reads represented 3.6 × 10−5–1.3% of the total sequencing reads (or 0.3–13 106 RPM) for each of the 22 RNA pools. Antimora rostrata (Gadiformes) with high body mass (>2 kg) had the lowest viral abundance, while Zoarcidae spp. (Scorpaeniformes) in the low body mass group (<0.11 kg) had the highest total virome abundance.

Ross Sea sampling site and phylogeny of fish hosts sampled for this study. Map of Antarctica with the Ross Sea region fish hosts that were sampled for this study highlighted (left). Phylogeny of 11 fish species and family groups sampled for this study (middle). Branches are highlighted in purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Sequence library read depths for all 22 libraries (two per host group) in millions (right). Low body mass libraries are the first bars for each host, and high body mass is the second (underneath).
Figure 1.

Ross Sea sampling site and phylogeny of fish hosts sampled for this study. Map of Antarctica with the Ross Sea region fish hosts that were sampled for this study highlighted (left). Phylogeny of 11 fish species and family groups sampled for this study (middle). Branches are highlighted in purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Sequence library read depths for all 22 libraries (two per host group) in millions (right). Low body mass libraries are the first bars for each host, and high body mass is the second (underneath).

Antarctic fish viromes contained sequences from likely vertebrate-infecting viruses that could be assigned to one of 21 different viral orders or families (Fig. 2). All 11 host species carried viruses from at least one taxonomic category. The number of viral taxonomic groups per host species ranged from one to eight, with an average of 4.4 per host. Trematomus (Perciformes) and Muraenolepis (Gadiformes) species carried the most (eight viral families), whereas the Perciformes species C. dewitti and P. scotti carried the least (one each). Only 4 of the 22 libraries contained no detectable fish virus sequences (low body mass P. barsukovi and M. caml and high body mass C. dewitti and P. scotti). Furthermore, transcripts from a total of 42 different viral polymerases or species were found across 15 RNA families, four DNA families, and one reverse-transcribed (RT) RNA family (Fig. 2, Table 2). Perciformes carried 23 unique viruses, Gadiformes carried 15, and four were found in the Scorpaeniformes. Picornavirus sequences were the most commonly identified (in five host species), followed by astroviruses (in four host species). Also of note was that the viruses identified were genetically distinct from any previously known viruses, representing entirely novel viromes. All of the viruses identified in this study shared 27–89% amino acid identity at most across all of the recovered virus segments (Table 2).

Antarctic fish viruses span 21 viral families and orders. Presence-absence map (top, middle) shows the presence of vertebrate viruses from 21 taxonomic groups in each host species group (top, left). Black boxes indicate the presence of the viral family/order in that host; white boxes indicate absence. Relative abundances (%) of each viral taxonomy in a host are shown (top, right). Network map showing connections between hosts and viral taxonomies (bottom, left). Fish hosts are shown by an illustration and viral families are indicated by coloured spheres. The width of grey connecting edges indicates abundance (in RPM).
Figure 2.

Antarctic fish viruses span 21 viral families and orders. Presence-absence map (top, middle) shows the presence of vertebrate viruses from 21 taxonomic groups in each host species group (top, left). Black boxes indicate the presence of the viral family/order in that host; white boxes indicate absence. Relative abundances (%) of each viral taxonomy in a host are shown (top, right). Network map showing connections between hosts and viral taxonomies (bottom, left). Fish hosts are shown by an illustration and viral families are indicated by coloured spheres. The width of grey connecting edges indicates abundance (in RPM).

Table 2.

Putative virus polymerases identified in this study.

LibraryMethodVirus taxonomyVirus nameBLASTp hit (RdRp/DdDp)aPercent identity (%)GenBank accession
TLO_HRNA-seqAmarillovirales; FlaviviridaeScaly rockcod flavivirusCAI5760484.1 Featherfin cichlid hepacivirus58.21PP590684
MRL_HRNA-seqArticulavirales; AmnoonviridaeMuraenolepis amnoonvirusQKU37010.1 Lauta virus31.72PP590685
PIM_HRNA-seqArticulavirales; AmnoonviridaeSpot-less plunderfish amnoonvirusQPC41259.1 Flavolineata virus64.60PP590686
TRD_HRNA-seqArticulavirales; AmnoonviridaeSlender scalyhead amnoonvirusQKU37010.1 Lauta virus34.59PP590687
ANT_LRNA-seqArticulavirales; OrthomyxoviridaeBlue antimora orthomyxovirusAIE52098.1 Influenza D virus40.45PP590688
LVD_LRNA-seqArticulavirales; OrthomyxoviridaeEelpouts orthomyxovirusAVM87616.1 Wenling orthomyxo-like virus 149.68PP590689
TLO_LRNA-seqBlubervirales; HepadnaviridaeScaly rockcod hepatitis B virusAYU58612.1 Eastern sea garfish hepatitis B virus57.85PP590690
TRD_LRNA-seqBlubervirales; HepadnaviridaeSlender scalyhead hepatitis B virusQWY26510.1 Hepatitis B virus60.99PP590691
MRL_HRNA-seqBunyavirales; ArenaviridaeMuraenolepis arenavirusAVM87649.1 Wenling frogfish arenavirus 253.38PP590692
TLO_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 149.95PP590693
TRD_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 151.98PP590694
BVK_HRNA-seqBunyavirales; HantaviridaeStubbeard plunderfish hantavirusYP_010085032.1 Wenling minipizza batfish hantavirus51.98PP590695
WGR_HRNA-seqBunyavirales; HantaviridaeMacrourus hantavirusYP_010088064.1 Actinovirus bernense43.75PP590696
TLO_HRNA-seqBunyavirales; PeribunyaviridaeScaly rockcod peribunyavirusYP_010840272.1 Largemouth bass bunyavirus59.85PP590697
TRD_LRNA-seqBunyavirales; PeribunyaviridaeSlender scalyhead peribunyavirusYP_010840272.1 Largemouth bass bunyavirus62.71PP590698
MRL_LRNA-seqChitovirales; PoxviridaeMuraenolepis poxvirusYP_009162498.1 Salmon gill poxvirus72.89PP590699
BVK_HRNA-seqGhabrivirales; TotiviridaeStubbeard plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 138.93PP590700
PIM_HRNA-seqGhabrivirales; TotiviridaeSpot-less plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 146.24PP590701
LVD_LRNA-seqHepevirales; HepeviridaeEelpouts hepe-like virusUHS16724.1 Herring hepe-like virus58.71PP590702
SZT_LRNA-seqHepevirales; HepeviridaeSaddleback plunderfish hepe-like virusUHS16724.1 Herring hepe-like virus60.00PP590703
TLO_HRNA-seqMononegavirales; FiloviridaeScaly rockcod filovirusQGM12359.2 Fiwi virus83.78PP590704
TRD_LRNA-seqMononegavirales; FiloviridaeSlender scalyhead filovirusWBV79890.1 Loetschberg virus81.72PP590705
LVD_HRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 1YP_009094346.1 Scophthalmus maximus rhabdovirus28.37PP590706
LVD_LRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 2AYU58616.1 Eastern sea garfish-associated rhabdo-like virus60.40PP590707
MRL_HRNA-seqMononegavirales; RhabdoviridaeMuraenolepis rhabdovirusQEU52490.1 Viral haemorrhagic septicaemia virus87.66PP590708
TLO_HRNA-seqMononegavirales; RhabdoviridaeScaly rockcod rhabdovirusUHR49730.1 Guiyang nephotettix cincticeps rhabdovirus 126.55PP590709
ANT_HRNA-seqNidovirales; CoronaviridaeBlue antimora coronavirusDAZ90582.1 Blenny letovirus40.00PP590710
MRL_HRNA-seqNidovirales; CoronaviridaeMuraenolepis coronavirusYP_010799913.1 Pacific salmon nidovirus42.34PP590711
WGR_HRNA-seqNidovirales; CoronaviridaeMacrourus coronavirusDAZ90588.1 Mudhopper letovirus52.93PP590712
PIM_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26315.1 Neolamprologus buescheri nackednavirus44.02PP590713
TLO_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.34PP590714
TRD_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.26PP590715
TRD_LRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.66PP590716
ANT_LRNA-seqPiccovirales; ParvoviridaeBlue antimora parvovirusWAQ80628.1 Parvovirinae sp.33PP590717
CHW_LRNA-seqPicornavirales; PicornaviridaeIcefish picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus44.70PP590718
MRL_HRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 1AVM87441.1 Wenling rattails picornavirus64.41PP590719
MRL_LRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 2YP_010796394.1 Wenling triplecross lizardfish picornavirus70.47PP590720
TLO_HRNA-seqPicornavirales; PicornaviridaeScaly rockcod picornavirusYP_010796391.1 Wenling crossorhombus picornavirus59.63PP590721
TRD_HRNA-seqPicornavirales; PicornaviridaeSlender scalyhead picornavirusYP_010796391.1 Wenling crossorhombus picornavirus73.13PP590722
WGR_LRNA-seqPicornavirales; PicornaviridaeMacrourus picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus33.57PP590723
PIM_LRNA-seqPimascovirales; IridoviridaeSpot-less plunderfish iridovirusQCF47437.1 Erythrocytic necrosis virus52.97PP590724
MRL_HRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus61.22PP590725
MRL_LRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus56.66PP590726
PIM_LRNA-seqReovirales; SpinareoviridaeSpot-less plunderfish reovirusAND67142.1 Grass carp reovirus56.88PP590727
TRD_LRNA-seqReovirales; SpinareoviridaeSlender scalyhead reovirusAMR58956.1 Grass carp reovirus68.67PP590728
MRL_HRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.40PP590729
MRL_LRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.66PP590730
PIM_HRNA-seqStellavirales; AstroviridaeSpot-less plunderfish astrovirusWLJ60740.1 Retropinna astrovirus 164.47PP590731
SRR2912518TSA miningBunyavirales; ArenaviridaeCatostomus commersonii arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus42.82
SRR3184758TSA miningBunyavirales; ArenaviridaeChanna punctata arenavirusYP_009551555.1 Wenling frogfish arenavirus 184.81
SRR12526228TSA miningBunyavirales; ArenaviridaeCoregonus artedi arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus51.99
LibraryMethodVirus taxonomyVirus nameBLASTp hit (RdRp/DdDp)aPercent identity (%)GenBank accession
TLO_HRNA-seqAmarillovirales; FlaviviridaeScaly rockcod flavivirusCAI5760484.1 Featherfin cichlid hepacivirus58.21PP590684
MRL_HRNA-seqArticulavirales; AmnoonviridaeMuraenolepis amnoonvirusQKU37010.1 Lauta virus31.72PP590685
PIM_HRNA-seqArticulavirales; AmnoonviridaeSpot-less plunderfish amnoonvirusQPC41259.1 Flavolineata virus64.60PP590686
TRD_HRNA-seqArticulavirales; AmnoonviridaeSlender scalyhead amnoonvirusQKU37010.1 Lauta virus34.59PP590687
ANT_LRNA-seqArticulavirales; OrthomyxoviridaeBlue antimora orthomyxovirusAIE52098.1 Influenza D virus40.45PP590688
LVD_LRNA-seqArticulavirales; OrthomyxoviridaeEelpouts orthomyxovirusAVM87616.1 Wenling orthomyxo-like virus 149.68PP590689
TLO_LRNA-seqBlubervirales; HepadnaviridaeScaly rockcod hepatitis B virusAYU58612.1 Eastern sea garfish hepatitis B virus57.85PP590690
TRD_LRNA-seqBlubervirales; HepadnaviridaeSlender scalyhead hepatitis B virusQWY26510.1 Hepatitis B virus60.99PP590691
MRL_HRNA-seqBunyavirales; ArenaviridaeMuraenolepis arenavirusAVM87649.1 Wenling frogfish arenavirus 253.38PP590692
TLO_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 149.95PP590693
TRD_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 151.98PP590694
BVK_HRNA-seqBunyavirales; HantaviridaeStubbeard plunderfish hantavirusYP_010085032.1 Wenling minipizza batfish hantavirus51.98PP590695
WGR_HRNA-seqBunyavirales; HantaviridaeMacrourus hantavirusYP_010088064.1 Actinovirus bernense43.75PP590696
TLO_HRNA-seqBunyavirales; PeribunyaviridaeScaly rockcod peribunyavirusYP_010840272.1 Largemouth bass bunyavirus59.85PP590697
TRD_LRNA-seqBunyavirales; PeribunyaviridaeSlender scalyhead peribunyavirusYP_010840272.1 Largemouth bass bunyavirus62.71PP590698
MRL_LRNA-seqChitovirales; PoxviridaeMuraenolepis poxvirusYP_009162498.1 Salmon gill poxvirus72.89PP590699
BVK_HRNA-seqGhabrivirales; TotiviridaeStubbeard plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 138.93PP590700
PIM_HRNA-seqGhabrivirales; TotiviridaeSpot-less plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 146.24PP590701
LVD_LRNA-seqHepevirales; HepeviridaeEelpouts hepe-like virusUHS16724.1 Herring hepe-like virus58.71PP590702
SZT_LRNA-seqHepevirales; HepeviridaeSaddleback plunderfish hepe-like virusUHS16724.1 Herring hepe-like virus60.00PP590703
TLO_HRNA-seqMononegavirales; FiloviridaeScaly rockcod filovirusQGM12359.2 Fiwi virus83.78PP590704
TRD_LRNA-seqMononegavirales; FiloviridaeSlender scalyhead filovirusWBV79890.1 Loetschberg virus81.72PP590705
LVD_HRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 1YP_009094346.1 Scophthalmus maximus rhabdovirus28.37PP590706
LVD_LRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 2AYU58616.1 Eastern sea garfish-associated rhabdo-like virus60.40PP590707
MRL_HRNA-seqMononegavirales; RhabdoviridaeMuraenolepis rhabdovirusQEU52490.1 Viral haemorrhagic septicaemia virus87.66PP590708
TLO_HRNA-seqMononegavirales; RhabdoviridaeScaly rockcod rhabdovirusUHR49730.1 Guiyang nephotettix cincticeps rhabdovirus 126.55PP590709
ANT_HRNA-seqNidovirales; CoronaviridaeBlue antimora coronavirusDAZ90582.1 Blenny letovirus40.00PP590710
MRL_HRNA-seqNidovirales; CoronaviridaeMuraenolepis coronavirusYP_010799913.1 Pacific salmon nidovirus42.34PP590711
WGR_HRNA-seqNidovirales; CoronaviridaeMacrourus coronavirusDAZ90588.1 Mudhopper letovirus52.93PP590712
PIM_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26315.1 Neolamprologus buescheri nackednavirus44.02PP590713
TLO_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.34PP590714
TRD_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.26PP590715
TRD_LRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.66PP590716
ANT_LRNA-seqPiccovirales; ParvoviridaeBlue antimora parvovirusWAQ80628.1 Parvovirinae sp.33PP590717
CHW_LRNA-seqPicornavirales; PicornaviridaeIcefish picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus44.70PP590718
MRL_HRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 1AVM87441.1 Wenling rattails picornavirus64.41PP590719
MRL_LRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 2YP_010796394.1 Wenling triplecross lizardfish picornavirus70.47PP590720
TLO_HRNA-seqPicornavirales; PicornaviridaeScaly rockcod picornavirusYP_010796391.1 Wenling crossorhombus picornavirus59.63PP590721
TRD_HRNA-seqPicornavirales; PicornaviridaeSlender scalyhead picornavirusYP_010796391.1 Wenling crossorhombus picornavirus73.13PP590722
WGR_LRNA-seqPicornavirales; PicornaviridaeMacrourus picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus33.57PP590723
PIM_LRNA-seqPimascovirales; IridoviridaeSpot-less plunderfish iridovirusQCF47437.1 Erythrocytic necrosis virus52.97PP590724
MRL_HRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus61.22PP590725
MRL_LRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus56.66PP590726
PIM_LRNA-seqReovirales; SpinareoviridaeSpot-less plunderfish reovirusAND67142.1 Grass carp reovirus56.88PP590727
TRD_LRNA-seqReovirales; SpinareoviridaeSlender scalyhead reovirusAMR58956.1 Grass carp reovirus68.67PP590728
MRL_HRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.40PP590729
MRL_LRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.66PP590730
PIM_HRNA-seqStellavirales; AstroviridaeSpot-less plunderfish astrovirusWLJ60740.1 Retropinna astrovirus 164.47PP590731
SRR2912518TSA miningBunyavirales; ArenaviridaeCatostomus commersonii arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus42.82
SRR3184758TSA miningBunyavirales; ArenaviridaeChanna punctata arenavirusYP_009551555.1 Wenling frogfish arenavirus 184.81
SRR12526228TSA miningBunyavirales; ArenaviridaeCoregonus artedi arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus51.99
a

Full sequence information for all viruses is available on GitHub.

Table 2.

Putative virus polymerases identified in this study.

LibraryMethodVirus taxonomyVirus nameBLASTp hit (RdRp/DdDp)aPercent identity (%)GenBank accession
TLO_HRNA-seqAmarillovirales; FlaviviridaeScaly rockcod flavivirusCAI5760484.1 Featherfin cichlid hepacivirus58.21PP590684
MRL_HRNA-seqArticulavirales; AmnoonviridaeMuraenolepis amnoonvirusQKU37010.1 Lauta virus31.72PP590685
PIM_HRNA-seqArticulavirales; AmnoonviridaeSpot-less plunderfish amnoonvirusQPC41259.1 Flavolineata virus64.60PP590686
TRD_HRNA-seqArticulavirales; AmnoonviridaeSlender scalyhead amnoonvirusQKU37010.1 Lauta virus34.59PP590687
ANT_LRNA-seqArticulavirales; OrthomyxoviridaeBlue antimora orthomyxovirusAIE52098.1 Influenza D virus40.45PP590688
LVD_LRNA-seqArticulavirales; OrthomyxoviridaeEelpouts orthomyxovirusAVM87616.1 Wenling orthomyxo-like virus 149.68PP590689
TLO_LRNA-seqBlubervirales; HepadnaviridaeScaly rockcod hepatitis B virusAYU58612.1 Eastern sea garfish hepatitis B virus57.85PP590690
TRD_LRNA-seqBlubervirales; HepadnaviridaeSlender scalyhead hepatitis B virusQWY26510.1 Hepatitis B virus60.99PP590691
MRL_HRNA-seqBunyavirales; ArenaviridaeMuraenolepis arenavirusAVM87649.1 Wenling frogfish arenavirus 253.38PP590692
TLO_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 149.95PP590693
TRD_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 151.98PP590694
BVK_HRNA-seqBunyavirales; HantaviridaeStubbeard plunderfish hantavirusYP_010085032.1 Wenling minipizza batfish hantavirus51.98PP590695
WGR_HRNA-seqBunyavirales; HantaviridaeMacrourus hantavirusYP_010088064.1 Actinovirus bernense43.75PP590696
TLO_HRNA-seqBunyavirales; PeribunyaviridaeScaly rockcod peribunyavirusYP_010840272.1 Largemouth bass bunyavirus59.85PP590697
TRD_LRNA-seqBunyavirales; PeribunyaviridaeSlender scalyhead peribunyavirusYP_010840272.1 Largemouth bass bunyavirus62.71PP590698
MRL_LRNA-seqChitovirales; PoxviridaeMuraenolepis poxvirusYP_009162498.1 Salmon gill poxvirus72.89PP590699
BVK_HRNA-seqGhabrivirales; TotiviridaeStubbeard plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 138.93PP590700
PIM_HRNA-seqGhabrivirales; TotiviridaeSpot-less plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 146.24PP590701
LVD_LRNA-seqHepevirales; HepeviridaeEelpouts hepe-like virusUHS16724.1 Herring hepe-like virus58.71PP590702
SZT_LRNA-seqHepevirales; HepeviridaeSaddleback plunderfish hepe-like virusUHS16724.1 Herring hepe-like virus60.00PP590703
TLO_HRNA-seqMononegavirales; FiloviridaeScaly rockcod filovirusQGM12359.2 Fiwi virus83.78PP590704
TRD_LRNA-seqMononegavirales; FiloviridaeSlender scalyhead filovirusWBV79890.1 Loetschberg virus81.72PP590705
LVD_HRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 1YP_009094346.1 Scophthalmus maximus rhabdovirus28.37PP590706
LVD_LRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 2AYU58616.1 Eastern sea garfish-associated rhabdo-like virus60.40PP590707
MRL_HRNA-seqMononegavirales; RhabdoviridaeMuraenolepis rhabdovirusQEU52490.1 Viral haemorrhagic septicaemia virus87.66PP590708
TLO_HRNA-seqMononegavirales; RhabdoviridaeScaly rockcod rhabdovirusUHR49730.1 Guiyang nephotettix cincticeps rhabdovirus 126.55PP590709
ANT_HRNA-seqNidovirales; CoronaviridaeBlue antimora coronavirusDAZ90582.1 Blenny letovirus40.00PP590710
MRL_HRNA-seqNidovirales; CoronaviridaeMuraenolepis coronavirusYP_010799913.1 Pacific salmon nidovirus42.34PP590711
WGR_HRNA-seqNidovirales; CoronaviridaeMacrourus coronavirusDAZ90588.1 Mudhopper letovirus52.93PP590712
PIM_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26315.1 Neolamprologus buescheri nackednavirus44.02PP590713
TLO_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.34PP590714
TRD_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.26PP590715
TRD_LRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.66PP590716
ANT_LRNA-seqPiccovirales; ParvoviridaeBlue antimora parvovirusWAQ80628.1 Parvovirinae sp.33PP590717
CHW_LRNA-seqPicornavirales; PicornaviridaeIcefish picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus44.70PP590718
MRL_HRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 1AVM87441.1 Wenling rattails picornavirus64.41PP590719
MRL_LRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 2YP_010796394.1 Wenling triplecross lizardfish picornavirus70.47PP590720
TLO_HRNA-seqPicornavirales; PicornaviridaeScaly rockcod picornavirusYP_010796391.1 Wenling crossorhombus picornavirus59.63PP590721
TRD_HRNA-seqPicornavirales; PicornaviridaeSlender scalyhead picornavirusYP_010796391.1 Wenling crossorhombus picornavirus73.13PP590722
WGR_LRNA-seqPicornavirales; PicornaviridaeMacrourus picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus33.57PP590723
PIM_LRNA-seqPimascovirales; IridoviridaeSpot-less plunderfish iridovirusQCF47437.1 Erythrocytic necrosis virus52.97PP590724
MRL_HRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus61.22PP590725
MRL_LRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus56.66PP590726
PIM_LRNA-seqReovirales; SpinareoviridaeSpot-less plunderfish reovirusAND67142.1 Grass carp reovirus56.88PP590727
TRD_LRNA-seqReovirales; SpinareoviridaeSlender scalyhead reovirusAMR58956.1 Grass carp reovirus68.67PP590728
MRL_HRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.40PP590729
MRL_LRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.66PP590730
PIM_HRNA-seqStellavirales; AstroviridaeSpot-less plunderfish astrovirusWLJ60740.1 Retropinna astrovirus 164.47PP590731
SRR2912518TSA miningBunyavirales; ArenaviridaeCatostomus commersonii arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus42.82
SRR3184758TSA miningBunyavirales; ArenaviridaeChanna punctata arenavirusYP_009551555.1 Wenling frogfish arenavirus 184.81
SRR12526228TSA miningBunyavirales; ArenaviridaeCoregonus artedi arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus51.99
LibraryMethodVirus taxonomyVirus nameBLASTp hit (RdRp/DdDp)aPercent identity (%)GenBank accession
TLO_HRNA-seqAmarillovirales; FlaviviridaeScaly rockcod flavivirusCAI5760484.1 Featherfin cichlid hepacivirus58.21PP590684
MRL_HRNA-seqArticulavirales; AmnoonviridaeMuraenolepis amnoonvirusQKU37010.1 Lauta virus31.72PP590685
PIM_HRNA-seqArticulavirales; AmnoonviridaeSpot-less plunderfish amnoonvirusQPC41259.1 Flavolineata virus64.60PP590686
TRD_HRNA-seqArticulavirales; AmnoonviridaeSlender scalyhead amnoonvirusQKU37010.1 Lauta virus34.59PP590687
ANT_LRNA-seqArticulavirales; OrthomyxoviridaeBlue antimora orthomyxovirusAIE52098.1 Influenza D virus40.45PP590688
LVD_LRNA-seqArticulavirales; OrthomyxoviridaeEelpouts orthomyxovirusAVM87616.1 Wenling orthomyxo-like virus 149.68PP590689
TLO_LRNA-seqBlubervirales; HepadnaviridaeScaly rockcod hepatitis B virusAYU58612.1 Eastern sea garfish hepatitis B virus57.85PP590690
TRD_LRNA-seqBlubervirales; HepadnaviridaeSlender scalyhead hepatitis B virusQWY26510.1 Hepatitis B virus60.99PP590691
MRL_HRNA-seqBunyavirales; ArenaviridaeMuraenolepis arenavirusAVM87649.1 Wenling frogfish arenavirus 253.38PP590692
TLO_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 149.95PP590693
TRD_LRNA-seqBunyavirales; ArenaviridaeTrematomus arenavirusYP_010839954.1 Salmon pescarenavirus 151.98PP590694
BVK_HRNA-seqBunyavirales; HantaviridaeStubbeard plunderfish hantavirusYP_010085032.1 Wenling minipizza batfish hantavirus51.98PP590695
WGR_HRNA-seqBunyavirales; HantaviridaeMacrourus hantavirusYP_010088064.1 Actinovirus bernense43.75PP590696
TLO_HRNA-seqBunyavirales; PeribunyaviridaeScaly rockcod peribunyavirusYP_010840272.1 Largemouth bass bunyavirus59.85PP590697
TRD_LRNA-seqBunyavirales; PeribunyaviridaeSlender scalyhead peribunyavirusYP_010840272.1 Largemouth bass bunyavirus62.71PP590698
MRL_LRNA-seqChitovirales; PoxviridaeMuraenolepis poxvirusYP_009162498.1 Salmon gill poxvirus72.89PP590699
BVK_HRNA-seqGhabrivirales; TotiviridaeStubbeard plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 138.93PP590700
PIM_HRNA-seqGhabrivirales; TotiviridaeSpot-less plunderfish totivirusQXJ19457.1 Bluegill toti-like virus 146.24PP590701
LVD_LRNA-seqHepevirales; HepeviridaeEelpouts hepe-like virusUHS16724.1 Herring hepe-like virus58.71PP590702
SZT_LRNA-seqHepevirales; HepeviridaeSaddleback plunderfish hepe-like virusUHS16724.1 Herring hepe-like virus60.00PP590703
TLO_HRNA-seqMononegavirales; FiloviridaeScaly rockcod filovirusQGM12359.2 Fiwi virus83.78PP590704
TRD_LRNA-seqMononegavirales; FiloviridaeSlender scalyhead filovirusWBV79890.1 Loetschberg virus81.72PP590705
LVD_HRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 1YP_009094346.1 Scophthalmus maximus rhabdovirus28.37PP590706
LVD_LRNA-seqMononegavirales; RhabdoviridaeEelpouts rhabdovirus 2AYU58616.1 Eastern sea garfish-associated rhabdo-like virus60.40PP590707
MRL_HRNA-seqMononegavirales; RhabdoviridaeMuraenolepis rhabdovirusQEU52490.1 Viral haemorrhagic septicaemia virus87.66PP590708
TLO_HRNA-seqMononegavirales; RhabdoviridaeScaly rockcod rhabdovirusUHR49730.1 Guiyang nephotettix cincticeps rhabdovirus 126.55PP590709
ANT_HRNA-seqNidovirales; CoronaviridaeBlue antimora coronavirusDAZ90582.1 Blenny letovirus40.00PP590710
MRL_HRNA-seqNidovirales; CoronaviridaeMuraenolepis coronavirusYP_010799913.1 Pacific salmon nidovirus42.34PP590711
WGR_HRNA-seqNidovirales; CoronaviridaeMacrourus coronavirusDAZ90588.1 Mudhopper letovirus52.93PP590712
PIM_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26315.1 Neolamprologus buescheri nackednavirus44.02PP590713
TLO_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.34PP590714
TRD_HRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.26PP590715
TRD_LRNA-seqOrtervirales; NackednaviridaeRoss Sea Perciformes nackednavirusWLN26309.1 Eretmodus cyanostictus nackednavirus46.66PP590716
ANT_LRNA-seqPiccovirales; ParvoviridaeBlue antimora parvovirusWAQ80628.1 Parvovirinae sp.33PP590717
CHW_LRNA-seqPicornavirales; PicornaviridaeIcefish picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus44.70PP590718
MRL_HRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 1AVM87441.1 Wenling rattails picornavirus64.41PP590719
MRL_LRNA-seqPicornavirales; PicornaviridaeMuraenolepis picornavirus 2YP_010796394.1 Wenling triplecross lizardfish picornavirus70.47PP590720
TLO_HRNA-seqPicornavirales; PicornaviridaeScaly rockcod picornavirusYP_010796391.1 Wenling crossorhombus picornavirus59.63PP590721
TRD_HRNA-seqPicornavirales; PicornaviridaeSlender scalyhead picornavirusYP_010796391.1 Wenling crossorhombus picornavirus73.13PP590722
WGR_LRNA-seqPicornavirales; PicornaviridaeMacrourus picornavirusUVF58819.1 Neopomacentrus bankieri picornavirus33.57PP590723
PIM_LRNA-seqPimascovirales; IridoviridaeSpot-less plunderfish iridovirusQCF47437.1 Erythrocytic necrosis virus52.97PP590724
MRL_HRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus61.22PP590725
MRL_LRNA-seqReovirales; SedoreoviridaeMuraenolepis reovirusAVM87459.1 Wenling scaldfish reovirus56.66PP590726
PIM_LRNA-seqReovirales; SpinareoviridaeSpot-less plunderfish reovirusAND67142.1 Grass carp reovirus56.88PP590727
TRD_LRNA-seqReovirales; SpinareoviridaeSlender scalyhead reovirusAMR58956.1 Grass carp reovirus68.67PP590728
MRL_HRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.40PP590729
MRL_LRNA-seqStellavirales; AstroviridaeMuraenolepis astrovirusAVM87498.1 Wenling snipefish astrovirus57.66PP590730
PIM_HRNA-seqStellavirales; AstroviridaeSpot-less plunderfish astrovirusWLJ60740.1 Retropinna astrovirus 164.47PP590731
SRR2912518TSA miningBunyavirales; ArenaviridaeCatostomus commersonii arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus42.82
SRR3184758TSA miningBunyavirales; ArenaviridaeChanna punctata arenavirusYP_009551555.1 Wenling frogfish arenavirus 184.81
SRR12526228TSA miningBunyavirales; ArenaviridaeCoregonus artedi arenavirusYP_010840780.1 Inner Mongolia sediment arena-like virus51.99
a

Full sequence information for all viruses is available on GitHub.

Vertebrate viral abundances were also variable across both host species and virus (Fig. 2). Abundances ranged from 0.02 RPM to ∼316 RPM. Perciformes and Muraenolepis spp. carried the highest number of viruses at the highest abundances. For example, viruses from the Arenaviridae, Iridoviridae, and Nackednaviridae were some of the most abundant in Perciformes (>20 RPM, Fig. 2). Other host species typically carried fewer than four different viruses and at lower abundances (<5 RPM each). We also assessed the potential effects of host taxonomy and body mass grouping on virome composition. Neither body mass (P = .376) nor host order (P = .205) had a significant impact on virome composition, whereas the host genus did (P = .013), although this could be confounded by the low sample sizes within each genus (Supplementary Fig. S5).

Viral diversity in Antarctic fishes also closely mirrored that of fish hosts in other marine ecosystems. To compare this diversity across studies, we examined the number of virus taxonomic groups (representing virus families and orders) and unique viral sequences likely representing a species per host species alongside the total number of individuals sampled in recent virome studies of fishes (Fig. 3, Supplementary Table S1 and Fig. S1). Generally, an increase in the number of individuals or host groups sampled corresponded with a greater number of unique viral species or sequences detected (R2 = 0.92, P < .01 for individuals; R2 = 0.85, P = .01 for host groups) (Supplementary Fig. S1). However, we found no significant correlation between the number of individuals sampled per host and the number of virus taxonomic groups uncovered in this study (R2 = −0.06, P = .86). While in-depth comparisons between polar and tropical fish viromes could not be made due to fish virome studies remaining limited, we could assess average diversity across some of the more comprehensive virome studies done to date. In the Ross Sea fishes, we identified viruses from 21 viral taxonomic groups, higher than the average of ∼14 virus families/orders across recent studies, and all of which have previously been identified in fish (Supplementary Table S1, Fig. 3). Antarctic fish harboured an average of 1.9 virus taxonomic groups per host, comparable to the average across studies of 1.8 (range: 0.33–2.2), and this observation was similar for the number of virus species per host (3.8) compared to the average (3.1; range: 1.2–5.5). Considering the influence of sample size on the number of viruses detected, we also examined the number of viruses per individual sampled. The average for Antarctic fish was 0.17, consistent with the overall average in other studies (0.17; range: 0.14–0.27).

Viral taxonomic groups present in recent fish virome studies. Chart of viral families and orders (i.e. Reovirales) identified in fish hosts in this study (Ross Sea, blue) and other recent metatranscriptomic studies. Black and blue boxes indicate the identification of a virus group in that study.
Figure 3.

Viral taxonomic groups present in recent fish virome studies. Chart of viral families and orders (i.e. Reovirales) identified in fish hosts in this study (Ross Sea, blue) and other recent metatranscriptomic studies. Black and blue boxes indicate the identification of a virus group in that study.

Phylogenetic analysis of Antarctic fish RNA viromes

We inferred phylogenetic relationships of the novel RNA virus species using viral transcripts containing the highly conserved RdRp sequence. Overall, we identified 39 RdRps, representing 35 potentially unique species across 10 host species and all three host orders sampled (Figs 4 and 5).

Phylogenetic analysis of negative-sense ssRNA viruses. Maximum likelihood trees of RdRp sequences of viruses from eight RNA virus families. Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).
Figure 4.

Phylogenetic analysis of negative-sense ssRNA viruses. Maximum likelihood trees of RdRp sequences of viruses from eight RNA virus families. Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).

Phylogenetic analysis of positive-sense ssRNA and dsRNA viruses. Maximum likelihood trees of RdRp sequences of viruses from four +ssRNA and two dsRNA virus families. Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).
Figure 5.

Phylogenetic analysis of positive-sense ssRNA and dsRNA viruses. Maximum likelihood trees of RdRp sequences of viruses from four +ssRNA and two dsRNA virus families. Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).

The viruses spanned negative-sense single-stranded (ss)RNA families, including the Amnoonviridae (n = 3 viruses), Arenaviridae (n = 2), Astroviridae (n = 3), Filoviridae (n = 2), Hantaviridae (n = 2), Peribunyaviridae (n = 2), Rhabdoviridae (n = 4), and Orthomyxoviridae (n = 2) (Fig. 4); positive-sense ssRNA families, including the Coronaviridae (n = 3), Flaviviridae (n = 1), Hepeviridae (n = 2), and Picornaviridae (n = 6); and the double-stranded (ds)RNA families, including the Spinareoviridae (n = 2), Sedoreoviridae (n = 1), and Totiviridae (n = 2) (Fig. 5). All of these viruses fell into clades that contained previously known fish viruses (Figs 4 and 5). The Picornaviridae family had the largest expansion with six new species from five Perciformes and Gadiformes hosts being added. Notably, viruses from Perciformes were placed in 13 of the 15 families, excluding the Coronaviridae and Sedoreoviridae families.

Negative ssRNA viruses

Antarctic fish similarly harboured a number of -ssRNA viruses falling within the Arenaviridae, Filoviridae, and Hantaviridae that include viruses causing severe haemorrhagic diseases in a variety of hosts (Cobo 2016) and in the Orthomyxoviridae, which includes respiratory disease-causing influenza and influenza-like viruses (Parry et al. 2020). Two new arenaviruses from Ross Sea fish, for example, provisionally termed Muraenolepis arenavirus and Trematomous arenavirus, shared ∼50% amino acid sequence identity with viruses such as Salmon pescarenavirus 1 and 2. Salmon arenaviruses have previously been associated with haemorrhagic symptoms in wild salmon (Oncorhynchus tshawytscha) (Mordecai et al. 2019). We also found rhabdoviruses in all three host orders studied here. One new virus, Muraenolepis rhabdovirus, shared >87% amino acid identity with Viral haemorrhagic septicaemia virus, which causes a potentially devastating haemorrhagic disease and infects a large number of species from the Perciformes and Gadiformes (Escobar et al. 2018). Other novel rhabdoviruses in scaly rockcod and eelpouts were more divergent in sequence, sharing <30% amino acid identity with other fish viruses linked to similar diseases, such as Scophthalmus maximus rhabdovirus (YP_009094346.1) (Zhang et al. 2007).

Positive ssRNA viruses

Viruses from the Letovirinae subfamily of the Coronaviridae were identified in a number of Gadiformes. The Letovirinae is comprised of viruses from aquatic vertebrates. Of note was the phylogenetic relationship between Macrourus coronavirus, found here, and Kanakana letovirus from diseased New Zealand lamprey (Geotria australis) (Miller et al. 2021, Fig. 2). That these viruses formed sister groups in the Letovirinae phylogeny is indicative of cross-order transmission over an uncertain time scale. The diversity of picornaviruses was also expanded by viruses in both Gadiformes and Perciformes. Novel picornaviruses shared between 34–70% amino acid identity with viruses from other bony fish from China (Shi et al. 2018) and coral reef systems (Costa et al. 2023a), falling in previously established subfamilies such as the Kodimesavirinae and Paavivirinae.

dsRNA viruses

We identified dsRNA viruses from the order Reovirales and the family Totiviridae. Totiviruses in two Pogonophryne hosts were both most closely related (sharing 39–46% amino acid identity) to Bluegill toti-like viruses found in deceased Lepomis macrochirus (bluegill) from Minnesota, USA (Sandlund et al. 2021). Interestingly, +ssRNA and dsRNA viruses from Perciformes and Gadiformes often fell into distinct clades or families. For example, both of the novel reoviruses from Perciformes hosts clustered in the Spinareoviridae family and were most closely related to viruses from the Aquareovirus genus, including Grass carp reovirus, which causes severe haemorrhagic disease in aquaculturally important species in China (Chen et al. 2013). Muraenolepis reovirus (from a Gadiformes species), on the other hand, fell into the Sedoreoviridae family most closely related to a likely benign virus, Wenling scaldfish reovirus, found in fish in China (Shi et al. 2018).

Newly identified DNA and RT DNA viruses

Two dsDNA viruses, one ssDNA virus, and RT DNA viruses (i.e. Hepadnaviridae) were found in Perciformes and Gadiformes, all falling into fish- or ectothermic vertebrate-associated groups of viruses (Fig. 6).

DNA and RT-DNA virus families expanded in this study. Maximum likelihood trees of DdDp or Pol proteins from viruses in dsDNA and ssDNA viral families and RT-DNA viruses from the Hepadnaviridae and Nackednaviridae. The hepadnavirus phylogeny is based on a tree from (Lauber et al. 2017). Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).
Figure 6.

DNA and RT-DNA virus families expanded in this study. Maximum likelihood trees of DdDp or Pol proteins from viruses in dsDNA and ssDNA viral families and RT-DNA viruses from the Hepadnaviridae and Nackednaviridae. The hepadnavirus phylogeny is based on a tree from (Lauber et al. 2017). Previously known and novel fish viruses are highlighted in blue. Novel viruses identified in this study are indicated by an illustration of their proposed host and highlighted based on the taxonomic order of their host: purple for Gadiformes, orange for Scorpaeniformes, and pink for Perciformes. Trees are rooted at midpoints and UF-bootstrap node support values ≥95% are denoted by an asterisk (*).

Hepadnaviruses (RT partially dsDNA viruses) are responsible for hepatitis B infections in some mammals, although they have also been identified in fish (Dill et al. 2016). Nackednaviruses, on the other hand, are a more recently identified group of exogenous replicating viruses in fish that may have diverged from a common ancestor of hepatitis B viruses >400 mya and have been hypothesized to help maintain the persistence of hepatitis B virus infections (Lauber et al. 2017). Two related hepadnavirus Pol genes falling in the Metahepadnavirus genus were found in both Trematomus species sharing 58–61% amino acid identity with the homologous sequences from hepatitis B viruses in fish (AYU58612.1) and amphibians (QWY26510.1) (Fig. 6, Table 2). We also found a full 3447 nt genome of a likely exogenous novel nackednavirus (Supplementary Fig. S1) tentatively named Ross Sea Perciformes nackednavirus (Fig. 6). The nackednavirus was found in both Trematomus hosts as well as a Pogonophryne species (P. immaculata) and shared ∼47% amino acid identity with Eretmodus cyanostictus nackednavirus (WLN26309.1) from a cichlid fish metagenome (Costa et al. 2023b).

Two additional dsDNA viruses most closely related to disease-causing viruses of fish and other cold-blooded vertebrates were identified in the Iridoviridae and Poxviridae (Fig. 6). Spot-less plunderfish iridovirus had ∼53% amino acid identity with Erythrocytic necrosis virus (QCF47437.1), which causes abnormalities of red blood cells and has been associated with mass mortalities in fish (Hershberger et al. 2009). Mauraenolepis poxvirus shared 73% amino acid identity with Salmon gill poxvirus (YP_009162498.1), which causes a potentially fatal respiratory disease in Atlantic salmon (Tartor et al. 2022).

Finally, we identified a ssDNA virus in the Parvoviridae in A. rostrata that exhibited ∼30% amino acid identity with Parvovirinae sp. (e.g. WAQ80628.1) from fish intestinal and metagenome samples (Reuter et al. 2022, Xi et al. 2023).

Recent and historic cross-species transmission of viruses

Evidence of recent and likely historic host-jumping events was found between and within host orders, particularly in the Perciformes. We observed a probable instance of recent cross-species virus transmission within the Arenaviridae. Specifically, Trematomus arenavirus was identified in both T. loennbergii (scaly rockcod) and T. lepidorhinus (slender scalyhead) (Fig. 4). The virus was 99% identical at the nucleotide level between the two hosts and did not fall below the abundance threshold, which would be suggestive of index-hopping with abundances of 8.8 and 221.9 RPM. It also shared ∼52% amino acid identity with the RdRp (L segment) of Salmon pescarenavirus (YP_010839954.1), representing a novel arenavirus most likely belonging to the Antennavirus genus of fish-infecting arenavirus (Pontremoli et al. 2019). The aforementioned Ross Sea Perciformes nackednavirus was also present in three different members of the Perciformes and similarly appears to represent another more recent, and the most widespread, host-switching event. Examples of more historic host-jumping of viruses between Perciformes, particularly Trematomus species, were noted in the Peribunyaviridae, Picornaviridae, Spinareoviridae, Totiviridae, and Hepadanaviridae families due to the presence of closely related, but not identical viruses. Similar patterns of host-switching were not observed within the equally sampled Gadiformes order. Additionally, we did observe two likely examples of viruses jumping host orders, including that of amnoonviruses between Muraenolepis (Gadiformes) and slender scalyhead (Perciformes; Trematomus) and of hepeviruses between Eelpouts (Scorpaeniformes; Zoarcidae) and Saddleback plunderfish (Perciformes; Pogonophryne).

Split NP in Trematomous hosts expands genomic and structural diversity of the Arenaviridae

Of particular note was Trematomous arenavirus that possessed a novel genome and NP structure (Fig. 7, Supplementary Fig. S2). The virus shared >99% nucleotide identity between pools of two different species in the genus Trematomous: T. lepidorhinus (slender scalyhead) and T. loennbergii (scaly rockcod) at abundances of 8.8 and 221.9 RPM, respectively (Fig. 7c). Trematomus arenavirus possessed the typical three genome segments—containing the polymerase, glycoprotein, and NP, found in other fish arenaviruses from the Antennavirus genus (Fig. 7a) (Pontremoli et al. 2019). However, the NP sequence in the novel virus was split over two ORFs as opposed to the canonical single ORF (Fig. 7, Supplementary Fig. S3). The split appeared to occur in a less conserved region of the NP, dividing it into two coding sequences containing its constituent NP core domain in one 349 amino acid segment and exonuclease domain in another 314 amino acid segment (Fig. 7a and b). Since this anomaly was present in the virus genomes recovered from both host species, at an abundance of >0.1% in the library with a lower viral abundance (8.8 RPM), and with good read coverage across the full sequence (Supplementary Fig. S3B), the split was not a result of a sequencing or assembly error nor a case of index-hopping. We note that the two coding regions were translated in different reading frames and a full, single protein encompassing both split ORFs can be resolved, for example, by insertion of a single base immediately before the stop codon of ORF1, suggesting a frameshift mutation as a possible cause for the split.

Trematomus arenavirus has a split NP. (a) Phylogeny (based on the RdRp sequence) of the Arenaviridae highlighted by genus (left). Trematomus arenavirus is highlighted in green and novel TSA-mined arenaviruses are highlighted in blue. Genome organizations of representative arenavirus genomes for each genus and their lengths (nt) showing the genomic structure diversity of the Arenaviridae (middle) and the protein structure of the segmented NP core (blue) and exonuclease domain (orange) of Trematomus arenavirus, as well as a superimposition of these on the Lassa virus NP structure (white) (Protein Data Bank: 3XM5), are shown (right). G = glycoprotein; N = nucleoprotein; P = polymerase (L segment); Z = Z protein; H = hypothetical protein. (b) Protein alignment of NPs from arenavirus species. The linker between the NP core and exonuclease domain is not well conserved across hosts. Note that the Trematomus arenavirus NP (shown in green) does not form a contiguous ORF and instead consists of two ORFs encoding different segments of the full NP (green). Trematomus arenavirus NP segments are highlighted in green and novel arenaviruses NPs identified from the TSA by this study are highlighted in dark blue. (c) Abundance (RPM) of host gene RPS13 (blue) compared to that of Trematomus arenavirus (green) in two Trematomus species.
Figure 7.

Trematomus arenavirus has a split NP. (a) Phylogeny (based on the RdRp sequence) of the Arenaviridae highlighted by genus (left). Trematomus arenavirus is highlighted in green and novel TSA-mined arenaviruses are highlighted in blue. Genome organizations of representative arenavirus genomes for each genus and their lengths (nt) showing the genomic structure diversity of the Arenaviridae (middle) and the protein structure of the segmented NP core (blue) and exonuclease domain (orange) of Trematomus arenavirus, as well as a superimposition of these on the Lassa virus NP structure (white) (Protein Data Bank: 3XM5), are shown (right). G = glycoprotein; N = nucleoprotein; P = polymerase (L segment); Z = Z protein; H = hypothetical protein. (b) Protein alignment of NPs from arenavirus species. The linker between the NP core and exonuclease domain is not well conserved across hosts. Note that the Trematomus arenavirus NP (shown in green) does not form a contiguous ORF and instead consists of two ORFs encoding different segments of the full NP (green). Trematomus arenavirus NP segments are highlighted in green and novel arenaviruses NPs identified from the TSA by this study are highlighted in dark blue. (c) Abundance (RPM) of host gene RPS13 (blue) compared to that of Trematomus arenavirus (green) in two Trematomus species.

At the structural level, the two NP domains overlap closely with the NP of Lassa virus (3MX5) (Fig. 7a) with the exception of an ∼60 amino acid α-helix structure in the N-terminus of the exonuclease domain with a high predicted local-distance difference test score (Jumper et al. 2021) (>80) and which was not present in other NP structures of known arenaviruses (Supplementary Fig. S4).

To determine if there were other arenaviruses with similar unique genomic and structural variations, we employed an additional data mining approach to screen available transcriptome assemblies in the NCBI TSA database. Using polymerase, glycoprotein, and NP sequences from Trematomous arenavirus as bait, we revealed an additional three fish-associated arenaviruses in C. punctata (SRR3184758), C. artedi (SRR12526228), and Castomus commersonii (SRR2912518) (Table 2, Fig. 7a and b, Supplementary Fig. S4). Although partial segments could be recovered for all three genomic segments of the three viruses, the viruses were not the closest relatives of Trematomous arenavirus (Fig. 7a) nor were the NP sequences similarly disrupted. Partial NPs of the viruses from C. comersonii and C. artedi did reveal similar helices linking the NP core and exonuclease domains (Fig. 7b, Supplementary Fig. S4), providing evidence for this helix structure in the NP as a unique feature of some aquatic arenaviruses.

Discussion

Climate and geographical changes can drive the evolution and adaptation of hosts and their viruses (Danovaro et al. 2011). Fish from the Southern Ocean experienced a significant reduction in biodiversity as a result of continental shifts and cooling water temperatures (Clarke and Johnston 1996, Eastman and Clarke 1998), but the results of these events on virome diversity in these fish have not been explored. To our knowledge, studies on the viruses of these fish have remained limited to instances of single virus genera or groups (e.g. bacteriophages) in Trematomus species (Van Doorslaer et al. 2018, Lopez et al. 2023). By analysing gill metatranscriptomes of 11 groups of fish collected from the Ross Sea region in Antarctica, we present entirely novel viromes that have greatly expanded the known aquatic vertebrate virosphere of Antarctica. In total, 42 unique fish viruses provide a new perspective on the virus diversity and evolution in these isolated, polar hosts.

Marine species richness significantly increases towards the equator (Rabosky et al. 2018). As the fishes of Antarctica are absent or rare in most other marine environments (Eastman 2005), our primary goal was to investigate whether the reduction in fish biodiversity during past continental and climatic changes would be reflected in their viromes compared to those in more temperate or tropical environments. Ross Sea fish viromes harboured an assortment of RNA and DNA viruses spanning 20 families. Despite the apparent novelty of their viromes, all of their viruses fell within phylogenetic groups of viruses that have commonly been identified in other fish virome-wide studies (Filipa-Silva et al. 2020, Costa et al. 2021, Geoghegan et al. 2021, Perry et al. 2022, Ford et al. 2024). Recent studies have also begun characterizing the viromes of fish in tropical ecosystems in more depth. The Great Barrier Reef in Australia, for example, boasts almost four times the number of fish species in the Southern Ocean (De’ath et al. 2012). Contrary to their limited host diversity, however, the observed viral family richness and the number and types of viral families infecting Ross Sea fishes were comparable to those of this coral reef system (Costa et al. 2023a, 2024). In fact, similar questions have also been asked of penguins and their ticks from the Antarctic Peninsula, revealing RNA virome diversity akin to that in Australian waterbirds (Wille et al. 2020). Overall, these emerging results suggest that Antarctica’s isolated nature and low community richness have not adversely affected the viral diversity of its fauna.

Animal radiations, such as those observed in notothenioid fishes like the Perciformes sampled here, likely played a role in shaping viral diversity in Antarctica. As host species that have undergone rapid speciation would be more genetically and ecologically similar, many of the immunological, cellular, and environmental barriers normally limiting microbial host-switching would yet to be established, in turn increasing the rate of cross-species virus transmission (Charleston et al. 2002, Longdon et al. 2014). As there is an inverse relationship between latitude and speciation in marine fish (Rabosky et al. 2018), it is possible that viral host-switching may be especially pronounced in marine environments with lower temperatures where endemism and speciation rates tend to be high. In the fish studied here, not only did the radiated Perciformes (notothenioid fish) harbour more viruses than the equally sampled Gadiformes, but the presence of closely related viruses, including those from the Arenaviridae, Nackednaviridae, and Spinareoviridae, shared across multiple notothenioid hosts, and the general connectivity of their shared family-level virome network, highlights the occurrence of numerous recent or historical viral host-switching incidents. A similar pattern is observed in the cichlid fish of Lake Tanganyika in Africa: the rapid radiation of these fish over the last 10 million years is associated with hosts that are very closely related and hence a high rate of cross-species virus transmission (Costa et al. 2024). This also illustrates that host evolutionary history, rather than environmental factors, is perhaps the most important determinant of rates of cross-species transmission and virus divergence.

In contrast to RNA viruses, on average DNA viruses tend to exhibit stronger congruency with their hosts’ evolutionary histories, indicative of long histories of co-divergence (Geoghegan et al. 2017, Peck et al. 2018). Accordingly, the identification of closely related hepadnaviruses in two Perciformes strengthens the argument that reduced host barriers to virus transmission have enabled host-jumping that has led to a higher number of closely related viral species persisting despite the speciation of notothenioids occurring ∼20 mya.

Continual global temperature changes are raising concerns about the imminent repercussions for biodiversity, species interactions, and infectious disease dynamics in aquatic species (Burge et al. 2014). Temperature is a key determinant in the distribution of ectohermic species and can impact interspecies interactions (Kordas et al. 2011) and disease susceptibility (Cohen et al. 2020, Páez et al. 2021). Rising water temperatures, for example, are forcing many aquatic species polewards (Nye et al. 2009, Danovaro et al. 2011). However, research addressing instances of viruses emerging as a direct consequence of altered host interactions in response to climate change remains limited. Although animals can host a range of viruses asymptomatically (Wille and Holmes 2020), the presence of viruses, like the novel iridovirus and poxvirus found here, closely related to known fish pathogens with high mortality rates (Nylund et al. 2008, Jung et al. 2017), should be of interest amid environmental shifts in regions with lower biodiversity, like the Southern Ocean (Duffy et al. 2016), where the risk of population declines due to emergent diseases is heightened by additional anthropogenic stressors such as overfishing and species introductions.

Trematomus arenavirus, found in scaly rockcod and slender scalyhead, also displayed a unique genomic variation. Unlike previously documented arenaviruses, Trematomus arenavirus had its NP sequence split over two non-overlapping ORFs. Arenavirus NPs are well conserved at both sequence and structural levels and are multifunctional, comprising a core domain that polymerizes and protects the viral RNA and an exonuclease domain for degrading dsRNA, connected by ‘flexible’ linkers (Papageorgiou et al. 2021). The split in the NP of Trematomus arenavirus occurred in the poorly conserved linker region between the two domains, which also contained an α-helix structure modelled with high confidence, uncommon in known arenaviruses and present in three of the four newly discovered fish arenaviruses. We provide evidence that this split may be the result of an indel mutation causing a frameshift. Gene fusion and fission—the joining or splitting of multiple genes/protein domains (Marsh and Teichmann 2010)—are fundamental evolutionary processes observed in cellular organisms (Snel et al. 2000). However, these processes have not been well studied in viruses. On the other hand, programmed ribosomal frameshifting, whereby ribosomes can be instructed to change reading frame via specific mRNA signals such as secondary RNA stem-loop structures, is a well-documented mechanism in RNA viruses such as HIV-1 and severe-acute-respiratory-syndrome-related coronavirus (Brierley and Dos Ramos 2006). Programmed ribosomal frameshifting allows viruses to encode multiple genes in overlapping or alternative reading frames, thereby optimizing their generally compact genomes. This implies that the novel virus may translate some combination of three possible proteins—the full NP through a readthrough of this frameshift, and the core and exonuclease domains separately. For now, the exact mechanisms behind this domain split and the well-structured helix, potentially illustrating adaptation to extreme temperatures affecting RNA secondary structures or protein folding (Feller 2018), remain unknown, as does any interaction between the two potentially independently functioning domains.

Our study has a number of limitations. Pooling samples into groups assists in the efficiency and likelihood of virus discovery as often a large number of individuals can be screened with low viral yield (Li et al. 2023), but this also limits conclusions that can be drawn due to often low resulting library (sample size) numbers. For instance, the capacity to estimate the prevalence of viruses in a host population or quantify the frequency of various evolutionary events is impacted. Tissue tropism is also frequently exhibited by viruses (Watashi and Wakita 2015), and thus the choice of tissue examined will necessarily influence what viruses can be uncovered. These viromes represent that of a single tissue—the gills. Nevertheless, gills are a common site of virus entry and infection in fish (Kim and Leong 1999) and analysis of viromes using multiple tissue types has shown that 94% of the viruses could be found in the gills alone (Costa et al. 2023b), so it is likely that a majority of the viruses infecting these Antarctic fish could be successfully identified here using gills.

In sum, we explored the implications of isolation and host adaptation in extreme environments on virus evolution by analysing the gill-associated viromes of fish from the Ross Sea, Antarctica. Our findings revealed an abundance of RNA and DNA viruses, greatly expanding the known virosphere of these polar species. Overall, their virome diversity mirrored that of fish from more diverse and warmer marine environments, suggestive of the maintenance of high viral diversity in these taxonomically restricted populations. While we provide a more extensive catalogue of the known viruses of Antarctic fishes, future research would benefit from broadening the scope to better understand the effects of host evolution and adaptation to polar environments on the evolution of their viruses or explore more directly whether viral diversity or virome connectivity differs greatly at the poles compared to temperate or tropical environments.

Acknowledgements

We acknowledge the crew of the San Aotea II, the New Zealand/Ministry for Primary Industries (MPI) observer Nigel Hollands, the CCAMLR international observer (CapFish) Chuma Sijaji, Mike Prasad (the scientist onboard), Jennifer Devine (the scientist onshore), and funding provided by Fisheries New Zealand under MPI project ANT2019-01C. We also thank Hamish Thompson for his Antarctic fish species illustrations.

Supplementary data

Supplementary data is available at VEVOLU online.

Conflict of interest:

None declared.

Funding

R.M.G. was funded by a University of Otago Doctoral Scholarship. J.L.G. is funded by a New Zealand Royal Society Rutherford Discovery Fellowship (RDF-20-UOO-007), a Marsden Fund Fast Start (20-UOO-105), and the New Zealand Ministry of Business, Innovation and Employment, Endeavour programme ‘Emerging Aquatic Diseases: a novel diagnostic pipeline and management framework’ (CAWX2207). E.C.H is funded by a National Health and Medical Research Council (Australia) Investigator Grant (GNT2017197).

References

Aguirre de Cárcer
D
,
López-Bueno
A
,
Alonso-Lobo
JM
et al.
Metagenomic analysis of lacustrine viral diversity along a latitudinal transect of the Antarctic Peninsula
.
FEMS Microbiol Ecol
2016
;
92
:fiw074.https://doi.org/10.1093/femsec/fiw074

Brierley
I
,
Dos Ramos
FJ
.
Programmed ribosomal frameshifting in HIV-1 and the SARS–CoV
.
Virus Res
2006
;
119
:
29
42
.https://doi.org/10.1016/j.virusres.2005.10.008

Buchfink
B
,
Reuter
K
,
Drost
H-G
.
Sensitive protein alignments at tree-of-life scale using DIAMOND
.
Nat Methods
2021
;
18
:
366
68
.https://doi.org/10.1038/s41592-021-01101-x

Burge
CA
,
Mark Eakin
C
,
Friedman
CS
et al.
Climate change influences on marine infectious diseases: implications for management and society
.
Ann Rev Mar Sci
2014
;
6
:
249
77
.https://doi.org/10.1146/annurev-marine-010213-135029

Camacho
C
,
Coulouris
G
,
Avagyan
V
et al.
BLAST+: architecture and applications
.
BMC Bioinf
2009
;
10
:421.https://doi.org/10.1186/1471-2105-10-421

Capella-Gutiérrez
S
,
Silla-Martínez
JM
,
Gabaldón
T
.
trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses
.
Bioinformatics
2009
;
25
:
1972
73
.https://doi.org/10.1093/bioinformatics/btp348

Charleston
MA
,
Robertson
DL
,
Sanderson
M
.
Preferential host switching by primate lentiviruses can account for phylogenetic similarity with the primate phylogeny
.
Syst Biol
2002
;
51
:
528
35
.https://doi.org/10.1080/10635150290069940

Chen
C
,
Sun
X
,
Liao
L
et al.
Antigenic analysis of grass carp reovirus using single-chain variable fragment antibody against IgM from Ctenopharyngodon idella
.
Sci China Life Sci
2013
;
56
:
59
65
.https://doi.org/10.1007/s11427-012-4425-5

Clarke
A
,
Johnston
IA
.
Evolution and adaptive radiation of Antarctic fishes
.
Trends Ecol Evol
1996
;
11
:
212
18
.https://doi.org/10.1016/0169-5347(96)10029-X

Cobo
F
.
Viruses causing hemorrhagic fever. Safety laboratory procedures
.
Open Virol J
2016
;
10
:
1
9
.https://doi.org/10.2174/1874357901610010001

Cohen
JM
,
Sauer
EL
,
Santiago
O
et al.
Divergent impacts of warming weather on wildlife disease risk across climates
.
Science
2020
;
370
:eabb1702.https://doi.org/10.1126/science.abb1702

Collins
MA
,
Brickle
P
,
Brown
J
et al.
The Patagonian toothfish: biology, ecology and fishery
.
Adv Mar Biol
2010
;
58
:
227
300
.

Costa
VA
,
Bellwood
DR
,
Mifsud
JCO
et al.
Limited cross-species virus transmission in a spatially restricted coral reef fish community
.
Virus Evol
2023a
;
9
:vead011.https://doi.org/10.1093/ve/vead011

Costa
VA
,
Bellwood
DR
,
Mifsud
JCO
et al.
Limited transmission of microbial species among coral reef fishes from the Great Barrier Reef, Australia
.
Curr Biol
2024
;
34
:
1247
57.e3
.https://doi.org/10.1016/j.cub.2024.02.008

Costa
VA
,
Mifsud
JCO
,
Gilligan
D
et al.
Metagenomic sequencing reveals a lack of virus exchange between native and invasive freshwater fish across the Murray-Darling Basin, Australia
.
Virus Evol
2021
;
7
:veab034.https://doi.org/10.1093/ve/veab034

Costa
VA
,
Ronco
F
,
Mifsud
JCO
et al.
Host adaptive radiation is associated with rapid virus diversification and cross-species transmission in African cichlid fishes
.
BioRxiv
2023b
.

Cummings
VJ
,
Bowden
DA
,
Pinkerton
MH
et al.
Ross Sea benthic ecosystems: macro- and mega-faunal community patterns from a multi-environment survey
.
Front Mar Sci
2021
;
8
:629787.https://doi.org/10.3389/fmars.2021.629787

Daane
JM
,
Detrich
HW
.
Adaptations and diversity of Antarctic fishes: a genomic perspective
.
Annu Rev Anim Biosci
2022
;
10
:
39
62
.https://doi.org/10.1146/annurev-animal-081221-064325

Danovaro
R
,
Corinaldesi
C
,
Dell’Anno
A
et al.
Marine viruses and global climate change
.
FEMS Microbiol Rev
2011
;
35
:
993
1034
.https://doi.org/10.1111/j.1574-6976.2010.00258.x

De’ath
G
,
Fabricius
KE
,
Sweatman
H
et al.
The 27-year decline of coral cover on the Great Barrier Reef and its causes
.
Proc Natl Acad Sci
2012
;
109
:
17995
99
.https://doi.org/10.1073/pnas.1208909109

Devries
AL
.
Glycoproteins as biological antifreeze agents in Antarctic fishes
.
Science
1971
;
172
:
1152
55
.https://doi.org/10.1126/science.172.3988.1152

Dill
JA
,
Camus
AC
,
Leary
JH
et al.
Distinct viral lineages from fish and amphibians reveal the complex evolutionary history of hepadnaviruses
.
J Virol
2016
;
90
:
7920
33
.https://doi.org/10.1128/JVI.00832-16

Duffy
JE
,
Lefcheck
JS
,
Stuart-Smith
RD
et al.
Biodiversity enhances reef fish biomass and resistance to climate change
.
Proc Natl Acad Sci
2016
;
113
:
6230
35
.https://doi.org/10.1073/pnas.1524465113

Eastman
JT
.
The nature of the diversity of Antarctic fishes
.
Polar Biol
2005
;
28
:
93
107
.https://doi.org/10.1007/s00300-004-0667-4

Eastman
JT
, and
Clarke
A
. A comparison of adaptive radiations of Antarctic fish with those of NonAntarctic fish. In:
Fishes of Antarctica
.
Milano
:
Springer
,
1998
,
3
26
. https://doi.org/10.1007/978-88-470-2157-0

Eastman
JT
,
Hubold
G
.
The fish fauna of the Ross Sea, Antarctica
.
Antarct Sci
1999
;
11
:
293
304
.https://doi.org/10.1017/S0954102099000383

Escobar
LE
,
Escobar-Dodero
J
,
Phelps
NBD
.
Infectious disease in fish: global risk of viral hemorrhagic septicemia virus
.
Rev Fish Biol Fish
2018
;
28
:
637
55
.https://doi.org/10.1007/s11160-018-9524-3

Eskandari
A
,
Leow
TC
,
Rahman
MBA
et al.
Antifreeze proteins and their practical utilization in industry, medicine, and agriculture
.
Biomolecules
2020
;
10
:1649.https://doi.org/10.3390/biom10121649

Feller
G
.
Protein folding at extreme temperatures: current issues
.
Semin Cell Dev Biol
2018
;
84
:
129
37
.https://doi.org/10.1016/j.semcdb.2017.09.003

Filipa-Silva
A
,
Parreira
R
,
Martínez-Puchol
S
et al.
The unexplored virome of two Atlantic coast fish: contribution of next-generation sequencing to fish virology
.
Foods
2020
;
9
:1634.https://doi.org/10.3390/foods9111634

Ford
CE
,
Dunn
CD
,
Leis
EM
et al.
Five species of wild freshwater sport fish in Wisconsin, USA, reveal highly diverse viromes
.
Pathogens
2024
;
13
:150.https://doi.org/10.3390/pathogens13020150

Geoghegan
JL
,
Di Giallonardo
F
,
Wille
M
et al.
Virome composition in marine fish revealed by meta-transcriptomics
.
Virus Evol
2021
;
7
:veab005.https://doi.org/10.1093/ve/veab005

Geoghegan
JL
,
Duchêne
S
,
Holmes
EC
.
Comparative analysis estimates the relative frequencies of co-divergence and cross-species transmission within viral families
.
PLoS Pathogens
2017
;
13
:e1006215.https://doi.org/10.1371/journal.ppat.1006215

Gille
ST
.
Warming of the Southern Ocean since the 1950s
.
Science
2002
;
295
:
1275
77
.https://doi.org/10.1126/science.1065863

Gong
Z
,
Liang
Y
,
Wang
M
et al.
Viral diversity and its relationship with environmental factors at the surface and deep sea of Prydz Bay, Antarctica
.
Front Microbiol
2018
;
9
:2981.https://doi.org/10.3389/fmicb.2018.02981

Grabherr
MG
,
Haas
BJ
,
Yassour
M
et al.
Full-length transcriptome assembly from RNA-Seq data without a reference genome
.
Nat Biotechnol
2011
;
29
:
644
52
.https://doi.org/10.1038/nbt.1883

Guindon
S
,
Dufayard
J-F
,
Lefort
V
et al.
New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0
.
Syst Biol
2010
;
59
:
307
21
.https://doi.org/10.1093/sysbio/syq010

Hershberger
PK
,
Elder
NE
,
Grady
CA
et al.
Prevalence of viral Erythrocytic necrosis in pacific herring and epizootics in Skagit Bay, Puget Sound, Washington
.
J Aquatic Animal Health
2009
;
21
:
1
7
.https://doi.org/10.1577/H08-035.1

Hoang
DT
,
Chernomor
O
,
von Haeseler
A
et al.
UFBoot2: improving the ultrafast bootstrap approximation
.
Mol Biol Evol
2018
;
35
:
518
22
.https://doi.org/10.1093/molbev/msx281

Holm
L
.
Dali server: structural unification of protein families
.
Nucleic Acids Res
2022
;
50
:
W210
W215
.https://doi.org/10.1093/nar/gkac387

Jumper
J
,
Evans
R
,
Pritzel
A
et al.
Highly accurate protein structure prediction with AlphaFold
.
Nature
2021
;
596
:
583
89
.https://doi.org/10.1038/s41586-021-03819-2

Jung
M-H
,
Nikapitiya
C
,
Vinay
T-N
et al.
Rock bream iridovirus (RBIV) replication in rock bream (Oplegnathus fasciatus) exposed for different time periods to susceptible water temperatures
.
Fish Shellfish Immunol
2017
;
70
:
731
35
.https://doi.org/10.1016/j.fsi.2017.09.038

Katoh
K
,
Standley
DM
.
MAFFT multiple sequence alignment software version 7: improvements in performance and usability
.
Mol Biol Evol
2013
;
30
:
772
80
.https://doi.org/10.1093/molbev/mst010

Kim
CH
, and
Leong
J-A
. FISH VIRUSES. In:
Encyclopedia of Virology
.
Elsevier
,
1999
,
558
68
. https://doi.org/10.1006/rwvi.1999.0100

Kordas
RL
,
Harley
CDG
,
O’Connor
MI
.
Community ecology in a warming world: the influence of temperature on interspecific interactions in marine systems
.
J Exp Mar Biol Ecol
2011
;
400
:
218
26
.https://doi.org/10.1016/j.jembe.2011.02.029

Kraberger
S
,
Austin
C
,
Farkas
K
et al.
Discovery of novel fish papillomaviruses: from the Antarctic to the commercial fish market
.
Virology
2022
;
565
:
65
72
.https://doi.org/10.1016/j.virol.2021.10.007

Langmead
B
,
Salzberg
SL
.
Fast gapped-read alignment with Bowtie 2
.
Nat Methods
2012
;
9
:
357
59
.https://doi.org/10.1038/nmeth.1923

Lauber
C
,
Seitz
S
,
Mattei
S
et al.
Deciphering the origin and evolution of hepatitis B viruses by means of a family of non-enveloped fish viruses
.
Cell Host Microbe
2017
;
22
:
387
399.e6
.https://doi.org/10.1016/j.chom.2017.07.019

Lefkowitz
EJ
,
Dempsey
DM
,
Hendrickson
RC
et al.
Virus taxonomy: the database of the International Committee on Taxonomy of Viruses (ICTV)
.
Nucleic Acids Res
2018
;
46
:
D708
17
.https://doi.org/10.1093/nar/gkx932

Li
B
,
Dewey
CN
.
RSEM: accurate transcript quantification from RNA-seq data with or without a reference genome
.
BMC Bioinf
2011
;
12
:323.https://doi.org/10.1186/1471-2105-12-323

Li
D
,
Liu
C-M
,
Luo
R
et al.
MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph
.
Bioinformatics
2015
;
31
:
1674
76
.https://doi.org/10.1093/bioinformatics/btv033

Li
YQ
,
Ghafari
M
,
Holbrook
AJ
et al.
The evolutionary history of hepaciviruses
.
BioRxiv
2023
.

Littlepage
JL
.
Oceanographic investigations in McMurdo Sound, Antarctica
.
Antarct Res Ser
2013
;
2
:
1
38
. https://doi.org/10.1029/AR005p0001

Longdon
B
,
Brockhurst
MA
,
Russell
CA
et al.
The evolution and genetics of virus host shifts
.
PLoS Pathog
2014
;
10
:e1004395.https://doi.org/10.1371/journal.ppat.1004395

Lopez
JKM
,
Austin
C
,
Farkas
K
et al.
Diverse microvirus genomes identified in the stomach of a sharp-spined Notothen (Trematomus pennellii) from the Ross Sea (East Antarctica)
.
Microbiol Resour Announc
2023
;
12
:
e01233
-
22
.https://doi.org/10.1128/mra.01233-22

Marsh
JA
,
Teichmann
SA
.
How do proteins gain new domains?
Genome Biol
2010
;
11
:126.https://doi.org/10.1186/gb-2010-11-7-126

Miller
AK
,
Mifsud
JCO
,
Costa
VA
et al.
Slippery when wet: cross-species transmission of divergent coronaviruses in bony and jawless fish and the evolutionary history of the Coronaviridae
.
Virus Evol
2021
;
7
:veab050.https://doi.org/10.1093/ve/veab050

Mirdita
M
,
Schütze
K
,
Moriwaki
Y
et al.
ColabFold: making protein folding accessible to all
.
Nat Methods
2022
;
19
:
679
82
.https://doi.org/10.1038/s41592-022-01488-1

Mordecai
GJ
,
Miller
KM
,
Di Cicco
E
et al.
Endangered wild salmon infected by newly discovered viruses
.
eLife
2019
;
8
:e47615.https://doi.org/10.7554/eLife.47615

Nelson
JS
.
Fishes of the World
, 4th edn.
Hoboken, NJ
:
Wiley & Sons
,
2006
. https://doi.org/10.1002/9781119174844

Nguyen
L-T
,
Schmidt
HA
,
von Haeseler
A
et al.
IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies
.
Mol Biol Evol
2015
;
32
:
268
74
.https://doi.org/10.1093/molbev/msu300

Nye
J
,
Link
J
,
Hare
J
et al.
Changing spatial distribution of fish stocks in relation to climate and population size on the Northeast United States continental shelf
.
Mar Ecol Prog Ser
2009
;
393
:
111
29
.https://doi.org/10.3354/meps08220

Nylund
A
,
Watanabe
K
,
Nylund
S
et al.
Morphogenesis of salmonid gill poxvirus associated with proliferative gill disease in farmed Atlantic salmon (Salmo salar) in Norway
.
Arch Virol
2008
;
153
:
1299
309
.https://doi.org/10.1007/s00705-008-0117-7

Oksanen
J
,
Simpson
GL
,
Blanchet
FG
et al.
vegan: Community Ecology Package
.
R package version 2.6-2
.
2022
.

Páez
DJ
,
Powers
RL
,
Jia
P
et al.
Temperature variation and host immunity regulate viral persistence in a salmonid host
.
Pathogens
2021
;
10
:855.https://doi.org/10.3390/pathogens10070855

Papageorgiou
N
,
Vaitsopoulou
A
,
Diop
A
et al.
Observation of arenavirus nucleoprotein heptamer assembly
.
FEBS Open Bio
2021
;
11
:
1076
83
.https://doi.org/10.1002/2211-5463.13106

Parry
R
,
Wille
M
,
Turnbull
O
et al.
Divergent influenza-like viruses of amphibians and fish support an ancient evolutionary association
.
Viruses
2020
;
12
:1042.https://doi.org/10.3390/v12091042

Peck
KM
,
Lauring
AS
,
Sullivan
CS
.
Complexities of viral mutation rates
.
J Virol
2018
;
92
:
10
128
.https://doi.org/10.1128/JVI.01031-17

Perry
BJ
,
Darestani
MM
,
Ara
MG
et al.
Viromes of freshwater fish with lacustrine and diadromous life histories differ in composition
.
Viruses
2022
;
14
:257.https://doi.org/10.3390/v14020257

Pettersen
EF
,
Goddard
TD
,
Huang
CC
et al.
UCSF ChimeraX: structure visualization for researchers, educators, and developers
.
Protein Sci
2021
;
30
:
70
82
.https://doi.org/10.1002/pro.3943

Pontremoli
C
,
Forni
D
,
Sironi
M
.
Arenavirus genomics: novel insights into viral diversity, origin, and evolution
.
Curr Opin Virol
2019
;
34
:
18
28
.https://doi.org/10.1016/j.coviro.2018.11.001

Rabosky
DL
,
Chang
J
,
Title
PO
et al.
An inverse latitudinal gradient in speciation rate for marine fishes
.
Nature
2018
;
559
:
392
95
.https://doi.org/10.1038/s41586-018-0273-1

Reuter
G
,
Boros
Á
,
Mátics
R
et al.
A novel parvovirus (family Parvoviridae) in a freshwater fish, zander (Sander lucioperca)
.
Arch Virol
2022
;
167
:
1163
67
.https://doi.org/10.1007/s00705-022-05419-5

Robichaux
S
,
Lacour
N
,
Bagby
GJ
et al.
Validation of RPS13 as a reference gene for absolute quantification of SIV RNA in tissue of rhesus macaques
.
J Virol Methods
2016
;
236
:
245
51
.https://doi.org/10.1016/j.jviromet.2016.08.004

Sandlund
L
,
Mor
SK
,
Singh
VK
et al.
Comparative molecular characterization of novel and known piscine toti-like viruses
.
Viruses
2021
;
13
:1063.https://doi.org/10.3390/v13061063

Seehausen
O
.
African cichlid fish: a model system in adaptive radiation research
.
Proc R Soc B
2006
;
273
:
1987
98
.https://doi.org/10.1098/rspb.2006.3539

Shi
M
,
Lin
X-D
,
Chen
X
et al.
The evolutionary history of vertebrate RNA viruses
.
Nature
2018
;
556
:
197
202
.https://doi.org/10.1038/s41586-018-0012-7

Smeele
ZE
,
Ainley
DG
,
Varsani
A
.
Viruses associated with Antarctic wildlife: from serology based detection to identification of genomes using high throughput sequencing
.
Virus Res
2018
;
243
:
91
105
.https://doi.org/10.1016/j.virusres.2017.10.017

Smith
WO
Ainley
DG
Cattaneo‐Vietti
R
et al. The Ross Sea continental shelf: regional biogeochemical cycles, trophic interactions, and potential future changes. In:
AD
Rogers
,
NM
Johnston
,
EJ
Murphy
et al. (eds),
Antarctic Ecosystems
.
Hoboken, NJ
:
Wiley
,
2012
,
213
42
. https://doi.org/10.1002/9781444347241.ch7

Snel
B
,
Bork
P
,
Huynen
M
.
Genome evolution
.
Trends Genet
2000
;
16
:
9
11
.https://doi.org/10.1016/S0168-9525(99)01924-1

Tartor
H
,
Dahle
MK
,
Gulla
S
et al.
Emergence of Salmon gill poxvirus
.
Viruses
2022
;
14
:2701.https://doi.org/10.3390/v14122701

Thacker
CE
,
Shelley
JJ
,
McCraney
WT
et al.
Delayed adaptive radiation among New Zealand stream fishes: joint estimation of divergence time and trait evolution in a newly delineated island species flock
.
Syst Biol
2021
;
71
:
13
23
.https://doi.org/10.1093/sysbio/syab014

Van Doorslaer
K
,
Kraberger
S
,
Austin
C
et al.
Fish polyomaviruses belong to two distinct evolutionary lineages
.
J Gen Virol
2018
;
99
:
567
73
.https://doi.org/10.1099/jgv.0.001041

Varsani
A
,
Frankfurter
G
,
Stainton
D
et al.
Identification of a polyomavirus in Weddell seal (Leptonychotes weddellii) from the Ross Sea (Antarctica)
.
Arch Virol
2017
;
162
:
1403
07
.https://doi.org/10.1007/s00705-017-3239-y

Watashi
K
,
Wakita
T
.
Hepatitis B virus and hepatitis D virus entry, species specificity, and tissue tropism
.
Cold Spring Harb Perspect Med
2015
;
5
:a021378.https://doi.org/10.1101/cshperspect.a021378

Wickham
H
.
Ggplot2: Elegant Graphics for Data Analysis
.
Springer
,
2016
.

Wille
M
,
Harvey
E
,
Shi
M
et al.
Sustained RNA virome diversity in Antarctic penguins and their ticks
.
ISME J
2020
;
14
:
1768
82
.https://doi.org/10.1038/s41396-020-0643-1

Wille
M
,
Holmes
EC
.
Wild birds as reservoirs for diverse and abundant gamma- and deltacoronaviruses
.
FEMS Microbiol Rev
2020
;
44
:
631
44
.https://doi.org/10.1093/femsre/fuaa026

Xi
Y
,
Jiang
X
,
Xie
X
et al.
Viromics reveals the high diversity of viruses from fishes of the Tibet highland
.
Microbiol Spectr
2023
;
11
:
e00946
-
23
.https://doi.org/10.1128/spectrum.00946-23

Zhang
Q
,
Tao
J
,
Gui
L
et al.
Isolation and characterization of Scophthalmus maximus rhabdovirus
.
Dis Aquat Organ
2007
;
74
:
95
105
.https://doi.org/10.3354/dao074095

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.

Supplementary data