#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Re-assessing the diversity of negative strand RNA viruses in insects


Authors: Simon Käfer aff001;  Sofia Paraskevopoulou aff001;  Florian Zirkel aff003;  Nicolas Wieseke aff004;  Alexander Donath aff002;  Malte Petersen aff005;  Terry C. Jones aff001;  Shanlin Liu aff007;  Xin Zhou aff008;  Martin Middendorf aff004;  Sandra Junglen aff001;  Bernhard Misof aff002;  Christian Drosten aff001
Authors place of work: Institute of Virology, Charité-Universitätsmedizin Berlin, Corporate Member of Free University, Humboldt-University and Berlin Institute of Health, Berlin, Germany aff001;  Center for Molecular Biodiversity Research, Zoological Research Museum Alexander Koenig, Bonn, Germany aff002;  Institute of Virology, University of Bonn Medical Centre, Bonn, Germany aff003;  Swarm Intelligence and Complex Systems Group, Department of Computer Science, Leipzig University, Germany aff004;  Senckenberg Biodiversity and Climate Research Centre, Senckenberg Gesellschaft für Naturforschung, Frankfurt am Main, Germany aff005;  Center for Pathogen Evolution, Department of Zoology, University of Cambridge, Cambridge, United Kingdom aff006;  BGI-Shenzhen, China Beishan Industrial Zone, Shenzhen, Guangdong Province, China aff007;  College of Food Science and Nutritional Engineering, China Agricultural University, Beijing, China aff008;  Beijing Advanced Innovation Center for Food Nutrition and Human Health, China Agricultural University, Beijing, China aff009;  German Center for Infection Research (DZIF), associated partner site Charité, Berlin, Germany aff010
Published in the journal: Re-assessing the diversity of negative strand RNA viruses in insects. PLoS Pathog 15(12): e32767. doi:10.1371/journal.ppat.1008224
Category: Research Article
doi: https://doi.org/10.1371/journal.ppat.1008224

Summary

The spectrum of viruses in insects is important for subjects as diverse as public health, veterinary medicine, food production, and biodiversity conservation. The traditional interest in vector-borne diseases of humans and livestock has drawn the attention of virus studies to hematophagous insect species. However, these represent only a tiny fraction of the broad diversity of Hexapoda, the most speciose group of animals. Here, we systematically probed the diversity of negative strand RNA viruses in the largest and most representative collection of insect transcriptomes from samples representing all 34 extant orders of Hexapoda and 3 orders of Entognatha, as well as outgroups, altogether representing 1243 species. Based on profile hidden Markov models we detected 488 viral RNA-directed RNA polymerase (RdRp) sequences with similarity to negative strand RNA viruses. These were identified in members of 324 arthropod species. Selection for length, quality, and uniqueness left 234 sequences for analyses, showing similarity to genomes of viruses classified in Bunyavirales (n = 86), Articulavirales (n = 54), and several orders within Haploviricotina (n = 94). Coding-complete genomes or nearly-complete subgenomic assemblies were obtained in 61 cases. Based on phylogenetic topology and the availability of coding complete genomes we estimate that at least 20 novel viral genera in seven families need to be defined, only two of them monospecific. Seven additional viral clades emerge when adding sequences from the present study to formerly monospecific lineages, potentially requiring up to seven additional genera. One long sequence may indicate a novel family. For segmented viruses, cophylogenies between genome segments were generally improved by the inclusion of viruses from the present study, suggesting that in silico misassembly of segmented genomes is rare or absent. Contrary to previous assessments, significant virus-host codivergence was identified in major phylogenetic lineages based on two different approaches of codivergence analysis in a hypotheses testing framework. In spite of these additions to the known spectrum of viruses in insects, we caution that basing taxonomic decisions on genome information alone is challenging due to technical uncertainties, such as the inability to prove integrity of complete genome assemblies of segmented viruses.

Keywords:

Insects – Phylogenetics – Phylogenetic analysis – Taxonomy – Viral genomics – Animal phylogenetics – Invertebrate genomics – Viral taxonomy

Introduction

Negative strand RNA viruses contain major groups of pathogenic viruses that cause rabies, hemorrhagic fevers, respiratory infections, measles, as well as a large range of important diseases and economically important conditions in livestock and plants [14]. Our current knowledge of negative strand RNA viruses is biased by the interest in medical disciplines and provides an incomplete image when it comes to more fundamental questions in viral evolution, such as the contribution of codivergence in the formation of major viral genetic lineages. These questions can only be addressed by systematic studies of larger taxonomic units of viral hosts, corresponding to whole orders or classes of animals, which is complicated by the difficulty to establish representative sample collections. Samples utilized for viral diversity studies are often collected on an opportunistic basis or repurposed from other studies, resulting in imbalance in host species representation, uncertainty in host classification, and uncertain assignment of samples. This is especially true for studies of insects that show an enormous genetic and morphological diversity.

Insects are the most speciose group of animals. Their origin has been dated to the early Ordovician, 479 million years ago, a time that predates the formation of terrestrial ecosystems [5]. Insects engage in symbiotic and parasitic relationships with a multitude of plants and animals, and are a vital component of the diet of animals, potentially facilitating virus transmission. Nevertheless, research on insect viruses has been mainly driven by interest in vector-borne diseases, resulting in virological studies that have focused on blood-feeding species, with rare exceptions [4, 6, 7]. However, blood-feeding insects represent only a minute fraction of the biological diversity of insects. Studies using massively parallel sequencing of collections of invertebrates have yielded an unprecedented diversity of novel RNA viruses [4, 6, 8]. However, the samples used in these studies only covered a limited range of insect species, contained many other groups of invertebrates such as spiders, worms, and molluscs, and were generated by sample pooling. Uncertain knowledge of host associations in these and other studies have caused a tendency to abandon host association as an important auxiliary criterion for taxonomic classification [9].

Here we systematically probed the diversity of negative strand RNA viruses in the largest and most representative collection of full transcriptome datasets of arthropods. The collection is designed to represent all extant lineages of Hexapoda without representational bias. This transcriptome database was first utilized for a phylogenomic re-assessment of the Hexapoda phylogeny in 2014, based on 103 full transcriptomes [9]. Since that time, the collection has been significantly extended to now cover 1243 full transcriptome datasets. All datasets including their corresponding unassigned contigs and scaffolds were screened for negative strand RNA viruses. The sample represents all orders of Insecta (insects, n = 1178), the orders Collembola (springtails, n = 23), Protura (coneheads, n = 4), and Diplura (n = 14) of Entognatha, as well as 24 outgroup species pertaining to Crustacea (n = 10), Myriapoda (n = 11), and Chelicerata (n = 3).

Results

We based our search on conserved sequence motifs within the RdRp gene that is present in the genomes of all replicating RNA viruses without a DNA stage except deltaviruses, and is not present in the genome of the eukaryotic or prokaryotic cell. We utilized profile hidden Markov models (pHMMs) to search for candidate viral RdRp motifs within 42,618,061 contigs and scaffolds which were 66 to 20,314 amino acids long. pHMMs were trained on template amino acid alignments covering the core conserved RdRp regions of representative viruses assigned to the families Rhabdoviridae, Paramyxoviridae, Filoviridae, Nyamiviridae, and Orthomyxoviridae, as well as the genera Orthonairovirus, Mammarenavirus, Jonvirus, Orthohantavirus, Orthobunyavirus, Tospovirus, Herbevirus, Phlebovirus, and Goukovirus.

According to the results of contig assembly, we initially detected 488 viral RdRp sequences. These were identified in 324 arthropod species belonging to all insect orders and several outgroup taxa. The host associations, exact taxonomic classification, as well as sampling sites of hosts for the viral genomes that appear in the phylogenetic trees in Fig 1 are summarized in S1 Table.

Maximum likelihood phylogenies of viruses found in the present study, viruses defined by ICTV as species, as well as selected unclassified viruses.
Fig. 1. Maximum likelihood phylogenies of viruses found in the present study, viruses defined by ICTV as species, as well as selected unclassified viruses.
Novel viruses from the present study are identified by an asterisk. Novel viruses to be considered for taxonomic proposals because of unique phylogenetic position and availability of a coding-complete genome are identified by numbers, corresponding to Table 1. Trees were inferred in RAxML based on alignments of viral sequences pertaining to: A: Rhabdoviridae; B: Xinmoviridae, Nyamiviridae, Bornaviridae, Artoviridae, Lispiviridae, Paramyxoviridae, Sunviridae, Filoviridae, and Pneumoniviridae; C: Chuviridae, Qinviridae, and Yueviridae; D: Orthomyxoviridae; E: Hantaviridae, Cruliviridae, Peribunyaviridae, and Fimoviridae; F: Phasmaviridae; G: Phenuiviridae; H: Arenaviridae, Mypoviridae, Nairoviridae, and Wupedeviridae. Branch colors show host associations: Black, viruses associated with vertebrates; red, insects; orange, invertebrates other than insects; green, plants. Bootstrap supports on tree nodes are shown by circles (> = 95) and triangles (>70). Designated host infraorders for Blattodea: CCR: Cockroaches; for Coleoptera: CCJ: Cucujiformia; for Diptera: CCM: Culicomorpha, MSM: Muscomorpha, PSM: Psychodomorpha; for Hemiptera: STR: Stenorrhyncha; for Hymenoptera: ACL: Aculeata, PRS: Parasitica; for Lepidoptera: HTN: Heteroneura; for Odonata: ANS: Anisoptera.

A large proportion of the viral sequences were co-detected with different pHMMs, owing to the nature of the search algorithm which makes the detection of distant homologs possible. This is of particular relevance since the template alignments, as well as the pHMM searches were done before the release of any of the sequences described in [4] and [8]. All data were later re-examined using BLASTp, with the inclusion of the data of [4] and [8], but no additional matches were retrieved. This confirms the sensitivity of the pHMM search approach and demonstrates that the search strategy is not biased by a virus reference library that stems from a fragmentary sample of insect species.

From the obtained contigs, 234 large sequences were selected for further analysis based on length, quality, and dissimilarity toward other sequences in the dataset. These sequences were later found to have highest similarity to members of Bunyavirales (n = 86), Articulavirales (n = 54), or Haploviricotina (n = 94), respectively. For non-segmented viruses, full genome assembly was often successful. For viruses with segmented genomes, assembly focused on the RdRp-encoding segment and was later on complemented by BLAST-based searches for other genome segments expected. Thereby, 218 coding sequences from genes that are not encoded on the same segment as the RdRp gene, such as glycoproteins, nucleoproteins, polymerase subunits, and proteins with unknown function were identified. Complete or coding-complete genomes were assembled in 61 cases. Many additional large but incomplete genome assemblies were obtained that contain open reading frames of unknown function and represent unknown genome architectures.

Of note, poly-A purification was applied on all samples due to the intended use for transcriptome analysis [5]. We expected a general loss of sensitivity in detecting viral sequences but did not expect a huge bias against viral genomes as opposed to viral mRNAs. For instance, even during single-cycle infection, more viral genomic RNA than viral mRNA is obtained from poly-A preparation in mononegaviruses, in spite their mRNAs but not genomes contain poly-A tails [10]. In natural infections an even larger excess of genomes over mRNA is expected, because genomic RNA accumulates once packed, while viral mRNA is degraded at a rate that weighs against new synthesis. Also in members of Bunyavirales a poly-A–related bias is not expected because neither mRNAs nor genomes of these viruses have poly-A tails.

Novel viruses found in this study are named after their host order, related viral family, and the designation “OKIAV” (for 1KITE insect-associated virus), followed by a number (e.g., Hemipteran orthomyxo-related virus OKIAV183). All sequences, including genome annotations, host associations, and sampling sites, are available on the Dryad Digital Repository under https://doi.org/10.5061/dryad.87vt6hm. All sequences will be released in GenBank along with the release of transcriptomes from the 1kite project (BioProject PRJNA183205: 'The 1KITE project: evolution of insects').

Phylogeny and implications on taxonomy

To enable phylogenetic analysis, contigs were translated and grouped into eight alignments based on preliminary sequence matching and phylogenies. The eight resulting trees as shown in Fig 1 were generated following intense optimization of alignments by trimming and focusing on conserved domains, with the aim to leave sufficient information in alignments while allowing large taxonomic groups of viruses to be compared. The trees A-H in Fig 1 cover the Rhabdoviridae (Fig 1A), all other Mononegavirales (Fig 1B), Chu-, Qin-, and Yueviridae (Fig 1C), Orthomyxoviridae (Fig 1D), Hanta-, Cruli-, Peribunya-, and Fimoviridae (Fig 1E), Phasmaviridae (Fig 1G), Phenuiviridae (Fig 1G), as well as Arena-, Mypo-, Nairo-, and Wupedeviridae (Fig 1H). Detailed phylogenies including host associations down to the host species level, as well as viral taxonomy information are shown in Supporting Information S1S3 and S5S33 Figs. A detailed description of current taxonomy including the novel virus findings, as well as classification suggestions resulting from the present data are provided in Supporting Information S1 Text.

The tree structures in Fig 1 suggest a remarkable separation between vertebrate and insect viruses, as noted already on the basis of a less inclusive sample of insect diversity [4, 8]. With the exception of the subtree of non-plant-associated rhabdoviruses that still remains star-like and may thus be undersampled, many insect-associated clades now appear well-differentiated with a balanced proportion of intermediate versus terminal branches. In spite of the more inclusive insect sampling contributed by the present study, novel insect viruses remain absent in well-known clades of pathogenic vertebrate viruses, such as the genus Lyssavirus, the families Paramyxoviridae, Bornaviridae, Filoviridae, Hantaviridae, and Arenaviridae. Also, some major groups of pathogenic arboviruses do not show an expansion of host associations following our search. For instance, the phleboviruses and orthobunyaviruses that are known to be mosquito-, sandfly-, midge-, or tick-borne, do not yield any novel insect-associated viruses in our sample in spite of its enormous genetic diversity (note that there are no sandfly, no mosquito, and only two midge species in our sample). This absence is remarkable as also the studies of [4] and [8] did not find any novel phleboviruses in the insects they sampled, while they did find novel phleboviruses in ticks. There may exist an ecologically-driven association of these viruses with blood-feeding insects. The additional association with ticks makes it possible that viruses could be exchanged between insects and ticks based on common bloodmeal sources.

Classification criteria exist only for a minority of viral genera. For instance, amino acid sequence distances of 4% and 5% have been proposed for species demarcation in orthobunyaviruses and phleboviruses, respectively. Our distance-based selection of sequences for inclusion in trees exceeds this distance criterion, making it likely that all novel sequences identified by an asterisk in Fig 1 could be classified as novel species. While many recently-described viruses still form solitary lineages in trees, a deep topological separation and structured host association emerges after inclusion of data from the present study. Large subclades within viral families are often associated with insect orders or suborders, indicating an important auxiliary criterion for subdivision of these viral families into genera. Structured host associations become particularly obvious in the families Rhabdoviridae, Xinmoviridae, Nyamiviridae, Artoviridae, Lispiviridae, Chuviridae, Phasmaviridae, and Feraviridae (Supporting Information S1 Text). Table 1 lists those virus clades that we propose to be considered for classification on the genus level based on sequence distance and host associations, while taking into account the completeness of available genome sequences. All in all, the analyzed sequences suggest a potential for classifying at least 27 novel genera based on coding-complete virus genomes, 20 of them without any previously known representative, and identify deep branching viral lineages that in the future may be classified as three novel families or subfamilies. Host associations are especially informative for subclassification of rhabdo- and xinmoviruses, chuviruses, orthomyxoviruses, and phasmaviruses. Furthermore, we add the first independent description of a full qinvirus genome (Collembolan qin-related virus OKIAV112), detected in the entognath Anurida maritima (seashore springtail, class Collembola) (Fig 1C). Detailed taxonomical considerations are provided in the Supporting Information S1 Text.

Tab. 1. List of phylogenetic groups to be considered for taxonomic proposals.
List of phylogenetic groups to be considered for taxonomic proposals.

The discovery of a large diversity of novel lineages warrants a re-assessment of the overall topology of negative strand viruses. Based on a manually curated alignment, we inferred a tree as shown in Fig 2 using Bayesian phylogeny. Only few of the topological relationships differ from those in Fig 1, which incorporates more alignment information specific for the smaller units of diversity covered therein. For instance, there is weak support for the branching point of rhabdoviruses, as also observed by [13]. As in Shi et al. [8] and as implied by current taxonomy, but unlike the results by Wolf et al. [13], the members of the order Articulovirales branch to the exclusion of all members of Bunyavirales, while the deep topology of Bunyavirales is well supported. A noteworthy finding is Collembolan phasma-related virus OKIAV223, a large sequence of 8154 nucleotides extending beyond the L-gene ORF, albeit not covering segment termini. It clusters with Phasmaviridae and branches from the phasmavirus lineage short after the split from the last common ancestor of Peribunyaviridae and Phasmaviridae (Fig 2). It is therefore the most appropriate outgroup for the peribunyavirus tree (Fig 1E), which has been considered for rooting that tree. It is interesting to note that this topology suggests acquisitions of tospo- and emaraviruses by plants from invertebrates, rather than an evolution of peribunyaviruses from plant viruses as suggested by alternative tree topologies. A number of other findings that indicate deeply diverged and novel virus groups are described in Supporting Information S1 Text.

Bayesian phylogeny of negative strand RNA viruses based on MrBayes.
Fig. 2. Bayesian phylogeny of negative strand RNA viruses based on MrBayes.
Red dots identify virus orders according to current taxonomy. Blue clouds show virus families. Selected viral species are identified for orientation. Asterisks indicate posterior probabilities > 0.9.

Genome composition

Tentative structures of complete or nearly complete genomes are summarized in Fig 3. It is noteworthy that genomes of chuviruses were found to appear in linear, circular, and segmented circular forms [4]. Fig 1C includes an additional 25 exemplary chuviruses from the present study, including seven with at least one full segment and one with two full segments. According to the mapping of raw RNAseq reads, all genomes or genome segments of these viruses are linear. Gene order is L-G-N, or N-G-L (Fig 3), confirming the two principal gene orders described in Li et al. [4]. Genomes with a missing glycoprotein gene or over-assembled contigs, as described in the same work, are not observed in the present study. To check for circular genome organization, we have re-mapped all raw RNAseq reads to consensus alignments of chuviral sequences joined head to tail. This approach did not find any reads crossing the potential head/tail sequence boundaries, as would be expected in the case of genome segments that are circular. While we do not claim to refute circular genomes in chuviruses, we cannot confirm this genome conformation based on our data (Supporting Information S34 Fig) and recommend further experimental validation.

Annotations of full or nearly full genomes and selected full bunyaviral L-segments found in the present study.
Fig. 3. Annotations of full or nearly full genomes and selected full bunyaviral L-segments found in the present study.
Annotation was done using InterProScan [14]. Bunyaviral putative nonstructural genes, as well as other significant subgenomic reading frames were annotated manually.

The genome segment termini in members of Bunyavirales form complementary panhandle structures [15, 16]. These short sequences are identical within, and similar between genome segments of a given viral genome, and are usually conserved in viruses that belong to one same genus. Because segment termini in most of the members of recently-defined novel viral genera have not been analyzed (including in the present study; refer to Supporting Information S1 Text), we determined segment co-segregation as an indicator of grouping congruence of genome elements. Tanglegrams are shown in in Supporting Information S15, S19, S24, S28, S29, and S33 Figs. In most major clades there is congruence among segments. Some clades, such as clade C of the orthomyxoviruses (Supporting Information S15 Fig) or the clades that define Shanga- and Herbevirus in the peribunyaviruses, show signs of reassortment in lineage precursors, as topological incongruence is observed for all members of the respective clades. In cases where individual incongruences are seen, such as in Dipteran phasma-related virus OKIAV224, Zorapteran phasma-related virus OKIAV242, or Coleopteran phasma-related virus OKIAV243, we cannot discriminate between in silico misassembly of genomes and actual reassortment based on the present data. Confirmation by virus isolation or re-sequencing including genome ends will be necessary.

To obtain an overall impression of segment co-segregation in newly-discovered segmented RNA viruses, we analyzed cophylogenies of RdRp-encoding segments and other segments from the same putative viral genomes using Jane [17]. We compared cophylogeny costs against that of datasets with randomized segment associations. As summarized in Fig 4, addition of the present findings rather improved the cophylogeny costs except in cophylogenies between L- and M-segments (RdRp- and glycoprotein-encoding) of phasma- and phenuiviruses where there was no relevant change (Fig 4). Also, in some viral trees the addition of the present data reveal segment cosegregation where this was not evident from the genomes of previously described viruses. We thus assume that our segmented genome findings overall do not suffer from in-silico misassembly or other artifacts. However, the genomes of exemplary strains defining novel genera should be confirmed experimentally.

Genome segment cophylogeny costs.
Fig. 4. Genome segment cophylogeny costs.
Red dotted lines show lowest possible costs for breaches of cophylogeny when using Jane [17] to optimize the tip opposition of RdRp-encoding genome segment trees with the corresponding trees for genome segments indicated in each figure panel name. Black curves show the same costs accrued during each of 1000 different randomizations of the dataset. Percent values in panels show the cophylogeny costs of the real cophylogeny dataset divided by the costs with randomization (medians from 1000 randomizations). Lower percent values indicate better agreement between the RdRp and the respective genome segment cophylogeny.

Host-virus co-segregation

Recent studies on invertebrate-associated RNA viruses found no evidence of host-virus co-segregation, and proposed frequent cross-host transmission of viruses between insect hosts that co-occupy the same ecological niches [4]. However, these results were based on a limited and spatially restricted sample of insects and other invertebrates. For the present study we have subjected all viral phylogenies to formal cophylogenetic comparisons on the basis of resolved and updated phylogenies of insects as in [5].

The phylogenies shown in Fig 1 were subjected to tests of breaches of cophylogeny using Jane [17] (refer to Supporting Information S1 Table for host associations). To determine the contribution of the novel sequences, separate analyses were done without the OKIAV sequences but incorporating all known and novel viruses as per ICTV taxonomy update end of 2018. The limited knowledge of host associations in most studies restricted the resolution of these analyses to the level of insect orders. Significant virus-host co-segregation was identified in both analyses (with and without OKIAV findings) for the majority of trees (Fig 5A). For the trees including Chuviridae and Hanta-/Peribunyaviridae, significant co-segregation could only be detected when including findings from the present study. Only for the tree summarizing members of Phasmaviridae, no host-virus co-segregation could be detected.

Analysis of host-virus phylogenetic co-segregation.
Fig. 5. Analysis of host-virus phylogenetic co-segregation.
A: Histograms of costs for co-segregation of randomized host associations (1000 iterations) compared to the original host associations (red dotted lines) for all phylogenies, without (left) and with (right) OKIAV sequence inclusion, calculated with Jane [17]. The p-values of all z-tests indicate an increase in costs of over 5% above the original costs. N.S. indicates non-significant cost differences. B: Ratio between number of co-speciations in original and randomized data sets calculated for Haploviricotina and Polyploviricotina with CoRe-PA [18]. C: Ratio between number of host switching events in original and randomized data sets calculated for Haploviricotina and Polyploviricotina with CoRe-PA.

Because analysis in Jane does not take branch lengths into account, does not incorporate topological uncertainty, and mainly penalizes breaches of phylogenetic congruence between virus and host tree, an alternative algorithm, CoRe-PA [18], was applied to further examine host-virus co-segregation. This more sophisticated algorithm takes into account the reconstruction of co-speciation and host-switching separately, and also discriminates these from duplication events. To exclude potential bias introduced by the analysis of small phylogenetic trees, we utilized trees that represent the complete tree topologies of Haploviricotina and Polyploviricotina, including topological uncertainty based on 1000 bootstrap replicates. All possible rootings (i.e., tree versions rooted on every branch) of all replicate trees were modeled in order to exclude the impact of false co-segregation signals among outgroup branches in insect and host trees. Randomization of host associations was performed as previously and used as null hypotheses in separate formal tests of co-segregation and host switching, respectively. As summarized in Fig 5B and 5C, trees of both virus subphyla showed significantly more co-segregation events with real host associations as opposed to randomized host associations. In contrast, tests for host-switching events were no different when analyzing real versus randomized host associations. This suggests that earlier indications of predominance of host-switching were likely to have been caused by uneven sampling, incorrect host attributions, or issues with viral phylogenies.

Discussion

Our study summarizes curated results from the most comprehensive metagenomic virus screen of a whole class of animals. The host dataset is resolved to all known orders of insects and involves an evenly selected representation of insect families. Because every sample has been genetically classified, the present work provides exact knowledge of host associations for a large range of insect-associated viruses for the first time. This knowledge is essential to study viruses as potential agents of vector-borne diseases in animals or plants, as well as to understand insect viruses in the context of biodiversity conservation and food production [19]. It also offers important criteria to define novel viral species and higher taxa. In this and other regards, our work complements the recent work by Li et al. [4] and Shi et al. [8] that was mainly focused on the description of viral diversity. For several taxa described in these studies, host associations are specified for the first time. Furthermore, some of the recent taxonomic proposals or re-classifications by ICTV are reconciled by the present data.

While our search has revealed an even greater diversity of positive strand RNA viruses, it has been challenging to curate large numbers of viral sequences while keeping track with taxonomy as it is being changed. The present frequency of taxonomy revisions makes it difficult to design basic analyses, such as phylogenies, so that they incorporate the latest viral taxonomy. The process of taxonomic classification owes to be more conservative and undergo slower cycles of renewal. It should allow time for independent reproduction of sequence findings before these go into taxonomic classification. Independent reproduction is becoming more important as ICTV has recently decided to allow classification on the sole basis of genome sequence data. As our study exemplifies, classification may be uncertain when full genomes cannot reliably be obtained, e.g., in segmented viruses. Genome segment termini should be tested for complementarity among segments, but are often not covered in sequencing studies [4, 8, 20]. Even in the few cases where genome ends are known and identical, this does not prove that detected segments stem from one viral genome, because the ends of e.g., bunyaviral segments are identical between viral species and can agree across genera and even families [21]. The study of viral sequences derived from mixtures of insects is particularly difficult, as they may be assembled into chimeric genomes whose fragments may, more likely, stem from different viruses than one virus. The ICTV executive board has stated, with regards to the danger of assembling chimeric genomes, that “These are all caveats that must be addressed experimentally for MG [metagenomic] sequence data to be used for classification purposes” [9]. From a virological view, segmented viruses that are to be classified as species based only on sequence data should have full genome coverage including genome ends, and should stem from individual insects. They should also be examined for phylogenetic congruence between genome segments. Multiple infections should be considered as a source of potential mis-classification. In the present study we were able to infer phylogenies of several segments other than RdRp-encoding segments and perform analyses on co-segregation between the RdRp and the proteins encoded by those segments. While many clades of accepted genera formed monophyletic groups in segment trees, the topology between trees was not always congruent (Supporting Information S15, S19, S24, S28, S29, and S33 Figs). This suggests that reassortment was involved in the formation of major viral taxa. Topological conflicts in individual virus genomes appeared to occur more often in clades consisting of recently discovered viruses ([4, 8, 20], and our study), than in known and functionally-characterized viruses. Even if the present data were not found to disturb topological congruence in phylogenetic analyses of segment association, this emphasizes the general concern regarding hasty classification of viruses identified in metagenomic studies in the absence of experimental evidence.

In many instances, the comprehensiveness of our dataset allows us to estimate the extent of genetic space occupied by taxonomic entities below the family level. Issues such as long-branch attraction and loss of information by reduction of alignments [22] are alleviated by the addition of novel members to formerly solitary lineages, even if their genome sequences are incomplete [23]. The delineation of taxon-specific genetic space is further assisted by reliance on the assumption of long-term co-segregation. In the absence of any other criteria, the knowledge of host associations can provide valuable information to define taxonomic units when newly discovered viruses are added to formerly monogeneric and monospecific families. In light of the obvious under-sampling within viral phylogenies, we regard it as an omission to not use host associations as assistance for virus classification decisions. Host associations allow the intuitive use of the principle of co-segregation that is confirmed for distinct, but related, insect-associated viruses in the present study. Several monophyletic clades in a number of viral families were found to be associated with defined orders of insects, yet all these associations were previously undiscovered [4, 8].

Our data suggest a potential to classify at least 27 novel genera (20 of them without any previously known species), and probably three novel families. Regarding viruses for which complete genomes or live isolates could not be obtained, the knowledge of host associations will orient future efforts to identify or isolate viruses in a more targeted manner. This will be of particular importance in insect groups that have relevance in food production or act as vectors of disease, as well as insect species that change their distribution and abundance due to environmental change.

The utilization of host associations for species classification is an accepted approach in virology, even if, in rare cases, spillovers or dual host associations will have to be taken into account and corrected for [24]. We expect cross-host transmission to be comparatively rare based on our analyses, and challenge the notion of frequent occurrence of cross host-virus transmission in the context of ecological and geographic proximity. The available studies are based on under-sampled viral trees and may suffer from other issues related to host assignment and topological correctness, in particular of viral phylogenies [4]. Tests of co-segregation need to take topological uncertainty into account and should be contrasted to null hypotheses.

We furthermore have concerns regarding the sole reliance on distance-based classification tools in the latest taxonomy proposals for negative strand RNA viruses [25, 26]. The utilized tool, DEmARC [27], analyzes concatenated sets of homologous protein domains that are conserved across the family of interest. It detects local regions of discontinuity in the pairwise distance spectrum within protein primary sequence alignments. Regions of discontinuity define putative limits of taxonomic units below the family level. It has only been evaluated for three families of vertebrate-infecting viruses (Picornaviridae, Coronaviridae, Filoviridae) and not for taxa that contain more than one family. Results for some of the larger virus groups, such as the genera Enterovirus versus Rhinovirus, have been controversial [28]. If applied to taxonomic units that exceed the family level, changes in genome architecture are common, and it becomes unclear whether protein-encoding genes other than the RdRp are homologous or have instead been acquired by lateral gene transfer [4, 8, 29]. The rate of recombination in newly-discovered viruses is unknown, and in segmented viruses this problem is aggravated by genome segment reassortment. Notably, DEmARC has not been validated at all for segmented viruses. On this basis, the analysis by distance-based tools has to be restricted to the RdRp gene and thus is hardly different from a phylogeny based only on RdRp genes. Other tools, such as the recently-proposed GRAViTy algorithm that incorporates genome composition, should be used and evaluated on the same problem [30].

As demonstrated in the present study and elsewhere [4, 8, 13], there is considerable topological uncertainty in large RNA virus phylogenies that comprise units of genetic diversity corresponding to viral orders. While genetic distance can be estimated in trees, its estimation is based on evolutionary models that cannot accommodate all biological factors of sequence evolution. For instance, viral population sizes, host generation times and infection or coinfection rates are expected to vary considerably between major lineages of viruses carried by hosts as different as mammals, birds, reptiles, fish, insects, spiders, crustaceans, worms, and protists. As these host associations may have been in place since millions of years, effects such as substitution saturation or footprints of recombination are expected to influence the inference of deep phylogenetic relationships. Considering the shortcomings of evolutionary distance as a single classification criterion, additional biological criteria including host associations should be included in taxonomic considerations, if only as a test of plausibility.

Like all other approaches to detect viruses based on sequence information, our work has several additional limitations. For instance, it remains difficult to differentiate endogenous viral elements from replicating viruses. Our approach to extract virus sequence data from insect transcriptomes makes it highly unlikely that integrated, non-transcribed viruses are falsely assigned as viral matches. Earlier studies have relied on transcript abundance of viral transcripts in comparison to total RNA content in RNAseq datasets, without the knowledge of whole transcriptome data. However, neither approach can exclude endogenous viral elements with certainty. For this reason, independent confirmation of findings is necessary before taxonomic conclusions are drawn. Our exact knowledge of host associations will enable host genome, as well as virus isolation studies that will ultimately exclude endogenization of viruses in the host germ line and confirm viral replication.

Materials and methods

Insect transcriptome data

We screened 1243 insect transcriptomes sequenced within the 1KITE project [5] including species from all recognized extant insect orders and additional arthropod orders. Samples were collected worldwide and RNAseq data were sequenced in an Illumina HiSeq2000 platform. Raw RNAseq data were assembled using SOAPdenovo-Trans-31kmer (version 1.01) [31], and checked for quality and cross contaminations with VecScreen (www.ncbi.nlm.nih.gov/tools/vecscreen/), and UniVec database build 7.0 (www.ncbi.nlm.nih.gov/tools/vecscreen/univec/).

Viral sequence generation and sorting

Template alignments for building profile hidden Markov models (pHMMs) were created using characterized RdRp amino acid sequences of negative strand RNA virus families. Due to the high sequence divergence of viral genes, even for closely related species, the sequence search was conducted at the amino acid level. Sequences were downloaded from the NCBI database, in October 2014, and aligned with the web-based alignment tool T-coffee in Expresso mode [32]. Transcriptome assemblies were translated in all six ORFs with the fastatranslate program within the package EXONERATE (version 2.2.0) [33]. This ORF library was scanned using HMMER version 3 [34] and only sequences with contiguous ORFs were regarded as viral matches. HMMER builds a pHMM from a template alignment and uses it to extract sequences that match the underlying probabilities of the model. This allows for detection of evolutionary distantly related sequences, with the advantage of remote RdRp homolog detection, but also the disadvantage of inflating the results with redundant duplicate sequences. Viral amino acid matches were checked for redundancy with a twofold approach: a) matches from each RdRp-pHMM were aligned to the original template alignments with MAFFT version 7.123, E-INS-i [35]. Poorly aligned regions and sequences that were too short and did not overlap with the selected alignment region were removed using trimAl [36] or manually upon inspection of the alignment in Geneious (Geneious v.9.1.8, Biomatters, Auckland, New Zealand, https://www.geneious.com), always complying with the preservation of known RdRp motifs. Trees were inferred with PhyML v.3.2.0 [37], using 1000 bootstrap replicates and Blosum62 amino acid substitution matrix. b) viral hits were compared with BLAST+ v.2.2.28+ [38] against the non-redundant NCBI protein database which had previously been filtered for viral sequences. This twofold approach enabled sorting the viral hits, thus removing any redundancy introduced by the HMM-search among different virus families.

Inference of alignments and phylogenies

To compose alignments for phylogeny, viral sequence hits (OKIAVs) from the present study were compared to GenBank as of August 2018 using BLASTP with an e-value cutoff of 10−6. Sequences over 30% similarity to any OKIAV were selected. All species and genera listed in the ICTV taxonomy table as of end of 2018 were added. During the revision, selected additional species from the ICTV species update released in February 2019 were added. The following literature contributions were additionally consulted and relevant unclassified viruses were added: [3957].

In total, 234 of the RdRp sequences found in insects in the present study were used for phylogenetic analyses (Supporting Information S1 Table). Alignments were calculated anew and refined with trimAl as described above. Model testing in MrBayes identified Blosum62 to be the amino acid substitution matrix compatible with all alignments. Trees were inferred in RAxML-NG version 0.7.0 BETA [58] plotting the transfer bootstrap expectation values [59]. Confirmatory phylogenetic analyses were done in PhyML v.3.2.0 [37] and MrBayes v3.2.6 [60], using the same substitution model and four different substitution rate categories with gamma distribution. For RAxML and PhyML, 1000 bootstrap replicates were computed, and for MrBayes chains were run until fully converged. All trees were plotted and annotated using the R package ggtree [61].

Virus genome organization

All ORFs of the full-length viral hits were annotated after comparing them against our customized viral database as well as with the InterProScan protein domain search tool [14].

Phylogenetic co-segregation of virus segments

Considering that orthomyxo- and bunyaviruses have segmented genomes, we additionally searched for proteins encoded by other segments (nucleoprotein, glycoprotein, PB2, and PA). For this search, we used the available protein sequences of the respective genera (NCBI) for a BLASTp search within the transcriptomes we had detected the RdRp segments in already. Trees shown in Supporting Information S15, S19, S24, S28, S29, and S33 Figs include only those taxa for which additional protein genes were found. The R package dendextend [62], was used to create tanglegram figures, that allow examination of topological consistency among the trees. Jane [17] was used to match trees of the RdRp-encoding genome segments to trees of the other segments, based on costs for breaches of cophylogeny (best match = lowest costs). Costs were also determined when segment-segment associations were randomized and these pairs of trees were then subjected to cophylogeny optimization in Jane. To obtain a quantitative measure of topological congruence, the costs associated with the real datasets were divided by the costs with randomization (median from 1000 randomizations). The resulting value is a percentage that expresses the cophylogeny cost relative to a randomly-associated cophylogeny of same tree size and structure (resulting in 0 for perfect cophylogeny and 1 for absence of cophylogeny). These relative costs are expressed as percentages in Fig 4. Because adding branches to cophylogenies must increase cophylogeny costs in randomized datasets, this was tested by Wilcoxon´s paired samples test. In all comparisons, the differences were highly significant.

Analysis of co-segregation of viruses with their hosts

Host-virus associations for each phylogenetic tree were examined to assess concurrent phylogenetic relationships using Jane [17]. As a basis for the host tree we used a modified version of the arthropod phylogeny from Misof et al. [5]. Since Jane needs a host for each taxon of the virus tree, unknown or undefined hosts cannot be assessed. Therefore, we added outlier branches of unidentified insect, unidentified arthropod, and non-arthropod hosts, to enable mapping to non-arthropod and pooled insect/arthropod hosts. The co-evolution costs of the original phylogenies were compared to 1000 iterations of randomized host-virus associations. A one-sided z-test (implemented in the R package BSDA [63]) was used to test whether the randomized costs are at least 5% higher than the original costs. This threshold was set to ensure that miniscule cost changes do not lead to false interpretations.

CoRe-PA [18], was used to evaluate the co-evolutionary dependencies of the two major virus subphyla of Haploviricotina and Polyploviricotina with their corresponding insect hosts. Given a co-segregating scenario, CoRe-PA aims to find the most parsimonious reconciliation between host and virus trees by evaluating four co-evolutionary events: co-speciation, sorting, duplication, and host-switching. Each type of event is assigned a certain cost and the co-phylogenetic assessment that minimizes the total cost of events is accepted. Since both Haploviricotina and Polyploviricotina trees were unrooted, all possible rooted versions were evaluated, meaning that for every edge a rooted tree was created. Insect phylogenies were rooted to the arthropod order Chelicerata. For each of the previously 1000 unrooted RAxML bootstrap trees, 229 and 266 rooted trees were created for Haploviricotina and Polyploviricotina respectively. A reconciliation for each of these trees to the corresponding insect phylogeny was computed. To determine the strength and significance of host-virus co-evolution, each reconciliation was compared against a randomized association of each co-phylogenetic scenario, keeping the tree topologies unchanged. 100 randomized scenaria were computed by randomly renaming the host tree tips. This preserved the structure of host-virus associations, while avoiding bias introduction from sampling random trees. To estimate the fit of each randomized scenario, reconciliations were computed with the following costs: 0 for sorting and duplication, -1000 for co-speciation, and -0.001 for a host-switching event.

Completeness of genome segments

The completeness of viral segments was assessed for all segmented-related findings. Segments with size similar to known relative viruses were regarded as at least complete-coding regions, if the segment ORF was terminated by a stop codon within the segment. Bunyaviruses have segments that form panhandles, with conserved, species-specific termini [15, 16]. We examined the genome termini for complementarity, and also evaluated whether the termini of one segment match those of the other segments.

Cytochrome oxidase subunit 1 (COI) barcode analysis

To investigate the possibility that other organisms were accidentally collected and therefore whether the hosts of the OKIAV viruses might not actually be the intended sampled organisms, a barcode search was conducted based on two databases. First, 2,534,455 cytochrome c oxidase subunit 1 (COI) gene sequences from GenBank sequences were retrieved on October 10, 2019 with the query “txid2759[Organism:exp] AND cytochrome oxidase subunit 1[All fields]”. Second, the German Barcode of Life (GBOL) database, which contains barcode sequences from species recovered in Germany (Animalia: 287,377 barcodes including 261,015 hexapods; Plantae 7,884 barcodes, Fungi 1,038 barcodes). Contigs assembled from the 1243 insect transcriptomes were matched against these databases using BLAST+ (version 2.6.0). The BLAST results were filtered for matches of length at least 500 nucleotides, with a nucleotide identity of at least 98%.

Of the 1243 insect transcriptomes, 34 (2.73%) have at least one contig that matches a non-Hexapoda barcode. The non-Hexapoda barcodes fall into 20 phylum/class categories, as shown in Supporting Information S3 Table.

Four of these 34 transcriptomes contained one or more negative strand RNA OKIAV, comprising a total of nine negative strand RNA viruses out of a total of 488 (1.8%) negative strand RNA viruses identified overall. Details of these four assemblies, their nine OKIAV viruses, and the matched non-Hexapoda barcodes are shown in Supporting Information S4 Table. Of the nine viruses, four are shown in the phylogenies in Fig 1, marked with an empty triangle to indicate the presence of a non-Hexapoda barcode in the associated assembly.

Supporting information

S1 Text [docx]
Supplementary results and discussion text.

S1 Table [xlsx]
Additional data for the identified viral genomes.

S2 Table [xlsx]
Read count and insect transcriptome library size for the full and nearly full OKIAV genomes.

S3 Table [xlsx]
Non-Hexapoda COI barcodes found across all insect transcriptome assemblies.

S4 Table [s]
Detection of non-Hexapoda COI barcodes.

S1 Fig [tif]
Viruses pertaining to .

S2 Fig [tif]
Maximum likelihood phylogenies with PhyML of viruses pertaining to .

S3 Fig [tif]
Bayesian phylogeny inference with MrBayes of viruses pertaining to .

S4 Fig [abtv]
Alignment of viroporins and the potential precursor ORF of Hymenopteran almendra-related virus OKIAV1.

S5 Fig [a]
Subtree of the sister clade of -, -, -, and .

S6 Fig [tif]
Viruses pertaining to , and .

S7 Fig [tif]
Maximum likelihood phylogeny with PhyML of viruses pertaining to , and .

S8 Fig [tif]
Bayesian phylogeny inference with MrBayes of viruses pertaining to , and .

S9 Fig [l]
Viruses pertaining to , , and .

S10 Fig [l]
Maximum likelihood phylogeny with PhyML of viruses pertaining to , , and .

S11 Fig [l]
Bayesian phylogeny inference with MrBayes of viruses pertaining to , , and .

S12 Fig [n]
Viruses pertaining to .

S13 Fig [n]
Maximum likelihood phylogeny with PhyML of viruses pertaining to .

S14 Fig [n]
Bayesian phylogeny inference with MrBayes of viruses pertaining to .

S15 Fig [tif]
Phylogenetic co-segregation between PB1 and PB2, PB1 and PA, and PB1 and nucleoprotein of the viruses pertaining to .

S16 Fig [tif]
Viruses pertaining to , and .

S17 Fig [n]
Maximum likelihood phylogeny with PhyML of viruses pertaining to , and .

S18 Fig [a]

S19 Fig [sub]
Phylogenetic co-segregation between RdRp and glycoprotein, and RdRp and nucleoprotein of the viruses pertaining to , and .

S20 Fig [tif]
Representative viruses of .

S21 Fig [n]
Viruses pertaining to .

S22 Fig [n]
Maximum likelihood phylogeny with PhyML of viruses pertaining to .

S23 Fig [n]
Bayesian phylogeny inference with MrBayes of viruses pertaining to .

S24 Fig [tif]
Phylogenetic co-segregation between RdRp and glycoprotein, and RdRp and nucleoprotein of the viruses pertaining to .

S25 Fig [n]
Viruses pertaining to .

S26 Fig [n]
Maximum likelihood phylogeny with PhyML of viruses pertaining to .

S27 Fig [n]
Bayesian phylogeny inference with MrBayes of viruses pertaining to .

S28 Fig [tif]
Phylogenetic co-segregation between RdRp and glycoprotein, and RdRp and nucleoprotein of the viruses pertaining to .

S29 Fig [tif]
Phylogenetic co-segregation between RdRp and glycoprotein, and RdRp and nucleoprotein of the viruses pertaining to and .

S30 Fig [n]
Viruses pertaining to , and .

S31 Fig [n]
Maximum likelihood phylogeny with PhyML of viruses pertaining to , and .

S32 Fig [n]
Bayesian phylogeny inference with MrBayes of viruses pertaining to , and .

S33 Fig [tif]
Phylogenetic co-segregation between RdRp and glycoprotein, and RdRp and nucleoprotein of the viruses pertaining to , and .

S34 Fig [g]
Read mapping on the genome of Odonatan chu-related virus OKIAV137.


Zdroje

1. Claas ECJ, Osterhaus AD, van Beek R, De Jong JC, Rimmelzwaan GF, Senne DA, et al. Human influenza A H5N1 virus related to a highly pathogenic avian influenza virus. Lancet. 1998;351(9101):472–7. doi: 10.1016/S0140-6736(97)11212-0 9482438

2. Drexler JF, Corman VM, Müller MA, Maganga GD, Vallo P, Binger T, et al. Bats host major mammalian paramyxoviruses. Nat Commun. 2012;3:796. doi: 10.1038/ncomms1796 22531181

3. Jones KE, Patel NG, Levy MA, Storeygard A, Balk D, Gittleman JL, et al. Global trends in emerging infectious diseases. Nature. 2008;451:990–3. doi: 10.1038/nature06536 18288193

4. Li C-X, Shi M, Tian J-H, Lin X-D, Kang Y-J, Chen L-J, et al. Unprecedented genomic diversity of RNA viruses in arthropods reveals the ancestry of negative-sense RNA viruses. Elife. 2015;4. doi: 10.7554/eLife.05378 25633976

5. Misof B, Liu S, Meusemann K, Peters RS, Donath A, Mayer C, et al. Phylogenomics resolves the timing and pattern of insect evolution. Science. 2014;346(6210):763–7. doi: 10.1126/science.1257570 25378627

6. Shi M, Lin X-D, Chen X, Tian J-H, Chen L-J, Li K, et al. The evolutionary history of vertebrate RNA viruses. Nature. 2018;556(7700):197–202. doi: 10.1038/s41586-018-0012-7 29618816

7. Junglen S, Drosten C. Virus discovery and recent insights into virus diversity in arthropods. Curr Opin Microbiol. 2013;16(4):507–13. doi: 10.1016/j.mib.2013.06.005 23850098

8. Shi M, Lin X-D, Tian J-H, Chen L-J, Chen X, Li C-X, et al. Redefining the invertebrate RNA virosphere. Nature. 2016;540(7634):539–43. doi: 10.1038/nature20167 27880757

9. Simmonds P, Adams MJ, Benkő M, Breitbart M, Brister JR, Carstens EB, et al. Virus taxonomy in the age of metagenomics. Nature Reviews Microbiology. 2017;15:161. doi: 10.1038/nrmicro.2016.177

10. Wignall-Fleming EB, Hughes DJ, Vattipally S, Modha S, Goodbourn S, Davison AJ, et al. Analysis of Paramyxovirus Transcription and Replication by High-Throughput Sequencing. J Virol. 2019;93(17). Epub 2019/06/14. doi: 10.1128/jvi.00571-19 31189700; PubMed Central PMCID: PMC6694822.

11. Di Giallonardo F, Audsley MD, Shi M, Young PR, McGraw EA, Holmes EC. Complete genome of Aedes aegypti anphevirus in the Aag2 mosquito cell line. Journal of General Virology. 2018;99(6):832–6. doi: 10.1099/jgv.0.001079 29741476

12. Schoonvaere K, De Smet L, Smagghe G, Vierstraete A, Braeckman BP, de Graaf DC. Unbiased RNA Shotgun Metagenomics in Social and Solitary Wild Bees Detects Associations with Eukaryote Parasites and New Viruses. PLoS One. 2016;11(12):e0168456. doi: 10.1371/journal.pone.0168456 28006002

13. Wolf YI, Kazlauskas D, Iranzo J, Lucia-Sanz A, Kuhn JH, Krupovic M, et al. Origins and Evolution of the Global RNA Virome. MBio. 2018;9(6). Epub 2018/11/30. doi: 10.1128/mBio.02329-18 30482837; PubMed Central PMCID: PMC6282212.

14. Jones P, Binns D, Chang H-Y, Fraser M, Li W, McAnulla C, et al. InterProScan 5: genome-scale protein function classification. Bioinformatics. 2014;30(9):1236–40. doi: 10.1093/bioinformatics/btu031 24451626

15. Elliott RM. Molecular biology of the Bunyaviridae. J Gen Virol. 1990;71 (Pt 3):501–22. doi: 10.1099/0022-1317-71-3-501 2179464

16. Mir MA, Brown B, Hjelle B, Duran WA, Panganiban AT. Hantavirus N protein exhibits genus-specific recognition of the viral RNA panhandle. J Virol. 2006;80(22):11283–92. doi: 10.1128/JVI.00820-06 16971445

17. Conow C, Fielder D, Ovadia Y, Libeskind-Hadas R. Jane: a new tool for the cophylogeny reconstruction problem. Algorithms Mol Biol. 2010;5:16. doi: 10.1186/1748-7188-5-16 20181081

18. Merkle D, Middendorf M, Wieseke N. A parameter-adaptive dynamic programming approach for inferring cophylogenies. BMC Bioinformatics. 2010;11 Suppl 1:S60. doi: 10.1186/1471-2105-11-S1-S60 20122236

19. Halloran A, Vantomme P, Hanboonsong Y, Ekesi S. Regulating edible insects: the challenge of addressing food security, nature conservation, and the erosion of traditional food culture. Food Security. 2015;7(3):739–46. doi: 10.1007/s12571-015-0463-8

20. Makhsous N, Shean RC, Droppers D, Guan J, Jerome KR, Greninger AL. Genome Sequences of Three Novel Bunyaviruses, Two Novel Rhabdoviruses, and One Novel Nyamivirus from Washington State Moths. Genome Announc. 2017;5(7). doi: 10.1128/genomeA.01668-16 28209840

21. Marklewitz M, Zirkel F, Rwego IB, Heidemann H, Trippner P, Kurth A, et al. Discovery of a unique novel clade of mosquito-associated bunyaviruses. J Virol. 2013;87(23):12850–65. doi: 10.1128/JVI.01862-13 24067954

22. Bergsten J. A review of long-branch attraction. Cladistics. 2005;21(2):163–93.

23. Wiens JJ. Can incomplete taxa rescue phylogenetic analyses from long-branch attraction? Syst Biol. 2005;54(5):731–42. doi: 10.1080/10635150500234583 16243761

24. Obbard DJ. Expansion of the metazoan virosphere: Progress, pitfalls, and prospects. Current opinion in virology. 2018;31:17–23. doi: 10.1016/j.coviro.2018.08.008 30237139

25. Maes P, Song T, Mark S, Paweska J, Song Q, Ye G, et al. ICTV taxonomic report 2017.016M.R.: Taxonomic expansion and reorganization of the order Mononegavirales.

26. Maes P, Alkhovsky S, Beer M, Briese T, Buchmeier MJ, Calisher CH, et al. ICTV taxonomic proposal 2017.012M: Taxonomic expansion and reorganization of the order Bunyavirales.

27. Lauber C, Gorbalenya AE. Partitioning the genetic diversity of a virus family: approach and evaluation through a case study of picornaviruses. J Virol. 2012;86(7):3890–904. doi: 10.1128/JVI.07173-11 22278230

28. Simmonds P. Methods for virus classification and the challenge of incorporating metagenomic sequence data. J Gen Virol. 2015;96(Pt 6):1193–206. doi: 10.1099/jgv.0.000016 26068186

29. Schuster S, Zirkel F, Kurth A, van Cleef KWR, Drosten C, van Rij RP, et al. A Unique Nodavirus with Novel Features: Mosinovirus Expresses Two Subgenomic RNAs, a Capsid Gene of Unknown Origin, and a Suppressor of the Antiviral RNA Interference Pathway. Journal of Virology. 2014;88(22):13447–59. doi: 10.1128/JVI.02144-14 25210176

30. Aiewsakun P, Adriaenssens EM, Lavigne R, Kropinski AM, Simmonds P. Evaluation of the genomic diversity of viruses infecting bacteria, archaea and eukaryotes using a common bioinformatic platform: steps towards a unified taxonomy. J Gen Virol. 2018;99:1331–43. doi: 10.1099/jgv.0.001110 30016225

31. Li R, Zhu H, Ruan J, Qian W, Fang X, Shi Z, et al. De novo assembly of human genomes with massively parallel short read sequencing. Genome Res. 2010;20(2):265–72. doi: 10.1101/gr.097261.109 20019144

32. Notredame C, Higgins DG, Heringa J. T-Coffee: A novel method for fast and accurate multiple sequence alignment. J Mol Biol. 2000;302(1):205–17. doi: 10.1006/jmbi.2000.4042 10964570

33. Slater GSC, Birney E. Automated generation of heuristics for biological sequence comparison. BMC Bioinformatics. 2005;6:31. doi: 10.1186/1471-2105-6-31 15713233

34. Eddy SR. Accelerated Profile HMM Searches. PLoS Comput Biol. 2011;7(10):e1002195. doi: 10.1371/journal.pcbi.1002195 22039361

35. Katoh K, Misawa K, Kuma K-I, Miyata T. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002;30(14):3059–66. doi: 10.1093/nar/gkf436 12136088

36. Capella-Gutierrez S, Silla-Martinez JM, Gabaldon T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics. 2009;25(15):1972–3. doi: 10.1093/bioinformatics/btp348 19505945

37. Guindon S, Dufayard J-F, Lefort V, Anisimova M, Hordijk W, Gascuel O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 2010;59(3):307–21. doi: 10.1093/sysbio/syq010 20525638

38. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, et al. BLAST+: architecture and applications. BMC Bioinformatics. 2009;10:421. doi: 10.1186/1471-2105-10-421 20003500

39. Contreras MA, Eastwood G, Guzman H, Popov V, Savit C, Uribe S, et al. Almendravirus: A Proposed New Genus of Rhabdoviruses Isolated from Mosquitoes in Tropical Regions of the Americas. Am J Trop Med Hyg. 2017;96:100–9. doi: 10.4269/ajtmh.16-0403 27799634

40. Reuter G, Boros Á, Pál J, Kapusinszky B, Delwart E, Pankovics P. Detection and genome analysis of a novel (dima)rhabdovirus (Riverside virus) from Ochlerotatus sp. mosquitoes in Central Europe. Infect Genet Evol. 2016;39:336–41. doi: 10.1016/j.meegid.2016.02.016 26883377

41. Shahhosseini N, Lühken R, Jöst H, Jansen S, Börstler J, Rieger T, et al. Detection and characterization of a novel rhabdovirus in Aedes cantans mosquitoes and evidence for a mosquito-associated new genus in the family Rhabdoviridae. Infect Genet Evol. 2017;55:260–8. doi: 10.1016/j.meegid.2017.09.026 28943405

42. Sun Q, Zhao Q, An X, Guo X, Zuo S, Zhang X, et al. Complete genome sequence of Menghai rhabdovirus, a novel mosquito-borne rhabdovirus from China. Arch Virol. 2017;162:1103–6. doi: 10.1007/s00705-016-3188-x 28000049

43. Tokarz R, Sameroff S, Tagliafierro T, Jain K, Williams SH, Cucura DM, et al. Identification of Novel Viruses in Amblyomma americanum, Dermacentor variabilis, and Ixodes scapularis Ticks. mSphere. 2018;3. doi: 10.1128/mSphere.00614-17 29564401

44. Xu CL, Cantalupo PG, Sáenz-Robles MT, Baldwin A, Fitzpatrick D, Norris DE, et al. Draft Genome Sequence of a Novel Rhabdovirus Isolated from Deinocerites Mosquitoes. Genome Announc. 2018;6. doi: 10.1128/genomeA.01438-17 29748415

45. Axén C, Hakhverdyan M, Boutrup TS, Blomkvist E, Ljunghager F, Alfjorden A, et al. Emergence of a new rhabdovirus associated with mass mortalities in eelpout (Zoarces viviparous) in the Baltic Sea. J Fish Dis. 2017;40:219–29. doi: 10.1111/jfd.12506 27416895

46. Charles J, Firth AE, Loroño-Pino MA, Garcia-Rejon JE, Farfan-Ale JA, Lipkin WI, et al. Merida virus, a putative novel rhabdovirus discovered in Culex and Ochlerotatus spp. mosquitoes in the Yucatan Peninsula of Mexico. J Gen Virol. 2016;97:977–87. doi: 10.1099/jgv.0.000424 26868915

47. Cholleti H, Hayer J, Abilio AP, Mulandane FC, Verner-Carlsson J, Falk KI, et al. Discovery of Novel Viruses in Mosquitoes from the Zambezi Valley of Mozambique. PLoS One. 2016;11:e0162751. doi: 10.1371/journal.pone.0162751 27682810

48. Lara Pinto AZd, Santos de Carvalho M, de Melo FL, Ribeiro ALM, Morais Ribeiro B, Dezengrini Slhessarenko R. Novel viruses in salivary glands of mosquitoes from sylvatic Cerrado, Midwestern Brazil. PLoS One. 2017;12:e0187429. doi: 10.1371/journal.pone.0187429 29117239

49. Økland AL, Nylund A, Øvergård A-C, Blindheim S, Watanabe K, Grotmol S, et al. Genomic characterization and phylogenetic position of two new species in Rhabdoviridae infecting the parasitic copepod, salmon louse (Lepeophtheirus salmonis). PLoS One. 2014;9:e112517. doi: 10.1371/journal.pone.0112517 25402203

50. Sabbadin F, Glover R, Stafford R, Rozado-Aguirre Z, Boonham N, Adams I, et al. Transcriptome sequencing identifies novel persistent viruses in herbicide resistant wild-grasses. Sci Rep. 2017;7:41987. doi: 10.1038/srep41987 28165016

51. Shi M, Neville P, Nicholson J, Eden J-S, Imrie A, Holmes EC. High-Resolution Metatranscriptomics Reveals the Ecological Dynamics of Mosquito-Associated RNA Viruses in Western Australia. J Virol. 2017;91. doi: 10.1128/JVI.00680-17 28637756

52. Simo Tchetgna HD, Nakoune E, Selekon B, Gessain A, Manuguerra J-C, Kazanji M, et al. Molecular Characterization of the Kamese Virus, an Unassigned Rhabdovirus, Isolated from Culex pruina in the Central African Republic. Vector Borne Zoonotic Dis. 2017;17:447–51. doi: 10.1089/vbz.2016.2068 28350284

53. Coffey LL, Page BL, Greninger AL, Herring BL, Russell RC, Doggett SL, et al. Enhanced arbovirus surveillance with deep sequencing: Identification of novel rhabdoviruses and bunyaviruses in Australian mosquitoes. Virology. 2014;448:146–58. doi: 10.1016/j.virol.2013.09.026 24314645

54. Contreras-Gutiérrez MA, Nunes MRT, Guzman H, Uribe S, Suaza Vasco JD, Cardoso JF, et al. Sinu virus, a novel and divergent orthomyxovirus related to members of the genus Thogotovirus isolated from mosquitoes in Colombia. Virology. 2017;501:166–75. doi: 10.1016/j.virol.2016.11.014 27936462

55. Ejiri H, Lim C-K, Isawa H, Fujita R, Murota K, Sato T, et al. Characterization of a novel thogotovirus isolated from Amblyomma testudinarium ticks in Ehime, Japan: A significant phylogenetic relationship to Bourbon virus. Virus Res. 2018;249:57–65. doi: 10.1016/j.virusres.2018.03.004 29548745

56. Medd NC, Fellous S, Waldron FM, Xuéreb A, Nakai M, Cross JV, et al. The virome of Drosophila suzukii, an invasive pest of soft fruit. Virus Evol. 2018;4:vey009. doi: 10.1093/ve/vey009 29644097

57. Rodrigues DSG, Medeiros DBdA, Rodrigues SG, Martins LC, de Lima CPS, de Oliveira LF, et al. Pacui Virus, Rio Preto da Eva Virus, and Tapirape Virus, Three Distinct Viruses within the Family Bunyaviridae. Genome Announc. 2014;2. doi: 10.1128/genomeA.00923-14 25395627

58. Kozlov AM, Darriba D, Flouri T, Morel B, Stamatakis A. RAxML-NG: A fast, scalable, and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics. 2019;35(21):4453–5. doi: 10.1093/bioinformatics/btz305 31070718

59. Lemoine F, B. Domelevo Entfellner J, Wilkinson E, Correia D, Dávila Felipe M, De Oliveira T, et al. Renewing Felsenstein’s phylogenetic bootstrap in the era of big data. Nature. 2018;556(7702):452–6. doi: 10.1038/s41586-018-0043-0 29670290

60. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, et al. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61(3):539–42. doi: 10.1093/sysbio/sys029 22357727

61. Yu G, Smith DK, Zhu H, Guan Y, Lam TT-Y. ggtree: anrpackage for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods Ecol Evol. 2016;8(1):28–36. doi: 10.1111/2041-210x.12628

62. Galili T. dendextend: an R package for visualizing, adjusting and comparing trees of hierarchical clustering. Bioinformatics. 2015;31(22):3718–20. doi: 10.1093/bioinformatics/btv428 26209431

63. Arnholt AT, Evans B. BSDA: Basic Statistics and Data Analysis. 2017.

Štítky
Hygiena a epidemiológia Infekčné lekárstvo Laboratórium

Článok vyšiel v časopise

PLOS Pathogens


2019 Číslo 12
Najčítanejšie tento týždeň
Najčítanejšie v tomto čísle
Kurzy

Zvýšte si kvalifikáciu online z pohodlia domova

Získaná hemofilie - Povědomí o nemoci a její diagnostika
nový kurz

Eozinofilní granulomatóza s polyangiitidou
Autori: doc. MUDr. Martina Doubková, Ph.D.

Všetky kurzy
Prihlásenie
Zabudnuté heslo

Zadajte e-mailovú adresu, s ktorou ste vytvárali účet. Budú Vám na ňu zasielané informácie k nastaveniu nového hesla.

Prihlásenie

Nemáte účet?  Registrujte sa

#ADS_BOTTOM_SCRIPTS#