J Gen VirolJ. Gen. VirolJGVvirThe Journal of General Virology0022-13171465-2099Society for General Microbiology24989172416593006842910.1099/vir.0.068429-0StandardAnimalRNA virusesConsensus proposals for classification of the family Hepeviridaehttp://vir.sgmjournals.orgClassification of the family HepeviridaeD. B. Smith and othersSmithDonald B.1SimmondsPeter1members of the International Committee on the Taxonomy of Viruses Hepeviridae Study GroupJameelShahid2EmersonSuzanne U.3HarrisonTim J.4MengXiang-Jin5OkamotoHiroaki6Van der PoelWim H. M.7PurdyMichael A.8University of Edinburgh, Centre for Immunity, Infection and Evolution, Edinburgh, Scotland, UKWellcome Trust/DBT India Alliance, Hyderabad, IndiaSpecial Volunteer, Retired Head Molecular Hepatitis Section, National Institute of Allergy and Infectious Diseases, National Institutes of Health, Bethesda, MD, USAUniversity College of London, London, UKCollege of Veterinary Medicine, Virginia Polytechnic Institute and State University, Blacksburg, VA, USADivision of Virology, Department of Infection and Immunity, Jichi Medical University School of Medicine, Tochigi-ken, JapanCentral Veterinary Institute, Wageningen University and Research Centre, Lelystad, The NetherlandsCenters for Disease Control and Prevention, National Center for HIV/Hepatitis/STD/TB Prevention, Division of Viral Hepatitis, Atlanta, GA, USACorrespondence Donald B. Smith D.B.Smith@ed.ac.uk10201410201495Pt 102223223218620143062014This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

The family Hepeviridae consists of positive-stranded RNA viruses that infect a wide range of mammalian species, as well as chickens and trout. A subset of these viruses infects humans and can cause a self-limiting acute hepatitis that may become chronic in immunosuppressed individuals. Current published descriptions of the taxonomical divisions within the family Hepeviridae are contradictory in relation to the assignment of species and genotypes. Through analysis of existing sequence information, we propose a taxonomic scheme in which the family is divided into the genera Orthohepevirus (all mammalian and avian hepatitis E virus (HEV) isolates) and Piscihepevirus (cutthroat trout virus). Species within the genus Orthohepevirus are designated Orthohepevirus A (isolates from human, pig, wild boar, deer, mongoose, rabbit and camel), Orthohepevirus B (isolates from chicken), Orthohepevirus C (isolates from rat, greater bandicoot, Asian musk shrew, ferret and mink) and Orthohepevirus D (isolates from bat). Proposals are also made for the designation of genotypes within the human and rat HEVs. This hierarchical system is congruent with hepevirus phylogeny, and the three classification levels (genus, species and genotype) are consistent with, and reflect discontinuities in the ranges of pairwise distances between amino acid sequences. Adoption of this system would include the avoidance of host names in taxonomic identifiers and provide a logical framework for the assignment of novel variants.

TheWellcome TrustOpenAccessEmbargo0
Introduction

Hepatitis E virus (HEV) is the cause of a self-limiting hepatitis with mortality rates of <2 % for immune competent individuals. However, higher mortality rates (10–30 %) are observed amongst pregnant women, while infection may become chronic in immunocompromised individuals. HEV was first recognized in the 1980s and its nucleotide sequence published in the 1990s (Huang et al., 1992; Reyes et al., 1990; Tam et al., 1991; Tsarev et al., 1992). Since then, HEV variants have been detected in a variety of human populations and in potential zoonotic sources such as pig, wild boar, deer, rabbit and mongoose. The variants that infect humans have been classified into four genotypes. Genotypes 1 and 2 are transmitted faecal–orally between humans, and genotypes 3 and 4 may be transmitted to humans zoonotically from infected pigs, deer and wild boar. These genotypes have been subdivided further into numerous subtypes (Lu et al., 2006), although the underlying criteria are controversial (Okamoto, 2007; Smith et al., 2013). A more distantly related virus detected in chickens also has been divided into genotypes (Bilic et al., 2009; Hsu & Tsai, 2014; Huang et al., 2004).

This relatively simple taxonomical landscape was disturbed recently by the discovery of viruses that are related to human HEV but infect rabbits (Zhao et al., 2009) and wild boar (Takahashi et al., 2011). Also, more divergent, HEV-like viruses have been described in rats (Johne et al., 2010a), ferrets (Raj et al., 2012) and bats (Drexler et al., 2012), and an even more divergent virus has been isolated from cutthroat trout (Batts et al., 2011). This last virus has a genome organization similar to that of HEVs, but shares very low levels of nucleotide and amino acid sequence identity. Finally, a virus detected in sewage has a partial genome sequence suggesting that it may represent a further, highly divergent member of the family Hepeviridae (Ng et al., 2012).

Several recent papers have attempted to summarize this diversity, but have unfortunately reached different conclusions, resulting in the use of multiple, contradictory definitions in the literature. For example, some authors have argued that avian HEV, bat HEV and cutthroat trout virus should be considered as belonging to different genera (Meng, 2013), while others have suggested that avian HEV, bat HEV and rat HEV should be considered as species within a single genus, with cutthroat trout virus as the sole member of a distinct genus (Smith et al., 2013). Both schemes are out-of-date in light of the recent publication of sequences from divergent variants isolated from moose (Lin et al., 2014), mink (Krog et al., 2013), fox (Bodewes et al., 2013), ferret (Li et al., 2014), wild boar (Takahashi et al., 2014) and camel (Woo et al., 2014).

The purpose of this paper is to present a consensus taxonomic framework that provides an agreed basis for the classification of currently described HEV variants, taking into account phylogenetic relationships, the extent of sequence identity and host range. This framework will facilitate the future classification of novel members of this family by providing researchers with straightforward guidelines for assigning new HEV variants. It also helps to clarify the zoonotic threat of particular HEV variants, so that, for example, rat HEV, which is a variant that has not so far been detected in humans, is clearly distinguished from variants of human HEV genotype 3 that are also found in rats (Lack et al., 2012). We propose the use of a common reference sequence and numbering system to simplify comparisons between different studies.

Our model for this consensus approach is that of Hepatitus C virus (HCV), in which the adoption of a consensus classification system (Simmonds et al., 2005) and numbering system with respect to a reference sequence (Kuiken et al., 2006) have stabilized the terminology used in HCV research and assisted researchers in providing unique and rational attribution of genotypes and subtypes. The utility of this framework is illustrated by the fact that an update to the HCV classification system almost 10 years later (Smith et al., 2014) was required simply to accommodate the large number of subsequently assigned subtypes, whereas only minor changes to the consensus rules for genotype and subtype assignment were necessary.

Results and DiscussionGenera

Members of the family Hepeviridae are positive-stranded RNA viruses with genomes of 6.6 to 7.3 kb. The longest ORF (ORF1) encodes a non-structural protein with several distinct domains: methyltransferase, Y-domain, papain-like protease, polyproline region (also known as the hypervariable region; HVR), macro domain, helicase and RNA-dependent RNA polymerase. ORF1 is followed by ORF2, which encodes the capsid protein, and ORF3, which overlaps with ORF2 and encodes a phosphoprotein that modulates cellular activities.

Despite this conserved genome structure, phylogenetic analysis within the family is complicated by difficulty in aligning the genome sequences of the most divergent variants. In particular, amino acid sequence identities between cutthroat trout virus and other members of the family are only 26–27 % (ORF1), 18–21 % (ORF2) and 13–16 % (ORF3). In contrast, identities between avian, rat and human HEV stand at 42–49 %, 42–55 % and 20–29 %, respectively (Batts et al., 2011). These average figures mask the fact that, in some genomic regions, no credible amino acid sequence alignment is achievable (Drexler et al., 2012; Huang et al., 2004; Johne et al., 2010b; Smith et al., 2013). Hence, phylogenetic comparisons simply based on pairwise (p)-distances between complete genome nucleotide sequences (Fig. 1a) may give a distorted impression of the relationships between variants. In addition, phylogenetic analysis must take into account the fact that substitutions at synonymous sites are saturated even in comparisons between the different genotypes of human HEV (Smith et al., 2013).

Phylogenetic analysis of members of the family Hepeviridae. (a) Neighbour-joining tree of p-distances among aligned complete genome sequences. (b) Maximum-likelihood tree for amino acid sequences in methyltransferase domain (ORF1-28 to ORF1-389) (c) helicase domain (ORF1-971 to ORF1-1185) (d) RNA-dependent RNA polymerase domain (ORF1-1249 to ORF-1-1671). (e) Proposed classification. Maximum-likelihood trees were computed using the model according to Le & Gascuel with a gamma distribution of evolutionary rates among sites with some invariant sites. Branches supported by >70 % of bootstrap replicates are indicated.

To recover more accurately sequence relationships between the most divergent variants in the Hepeviridae, we screened for regions of the genome that are clearly homologous using the Motif Scan program (http://myhits.isb-sib.ch). This identified three subgenomic regions in ORF1 comprising the methyltransferase [ORF1 residues 28 (ORF1-28) to ORF1-389 numbered relative to the sequence of the HEV Burma isolate, GenBank accession M73218], helicase (ORF1-971 to ORF1-1185) and RNA-dependent RNA polymerase (ORF1-1249 to ORF1-1671). Maximum-likelihood analysis of alignments of each region reproduced phylogenetic trees with a similar topology but with much shorter terminal branches (Fig. 1b–d) than obtained from complete genome sequences. Branch lengths in these trees likely represent better reconstructions of evolutionary depth and demonstrate that cutthroat trout virus is substantially more divergent from other members of the family Hepeviridae than indicated by nucleotide sequence comparisons. Similarly, p-distances amongst these subgenomic amino acid sequences form non-overlapping distributions, with the greatest distances observed in comparisons including cutthroat trout virus (Fig. 2). In addition, cutthroat trout virus has an aberrant genome organization (ORF3 is displaced towards the middle of ORF2) and a distinct host range (fish rather than mammals or birds). For these reasons, we recommend that cutthroat trout virus should be assigned to the new genus Piscihepevirus (Table 1), as proposed previously (Meng, 2013), with the single species Piscihepevirus A. We prefer these genus and species names to Cutrovirus (Batts et al., 2011), since the virus appears to be confined to fish, having also been detected in five other species of trout (Batts et al., 2011).

Frequency distribution of distances among the sequences of members of the family Hepeviridae. Plots show the frequency of amino acid sequence p-distances among amino acid sequences of the methyltransferase (ORF1-28 to ORF1-389, dotted line), helicase (ORF1-971 to ORF1-1185, solid line) and RdRP (ORF1-1249 to ORF1-1671, dashed line) domains. The methyltransferase and helicase distance frequencies are shifted by 160 and 80 respectively in order to improve legibility. Filled bars indicate the range of inter-generic distances.

Proposed classification of the family <italic>Hepeviridae</italic>
FamilyGenusSpeciesPrototype isolateGenBank accessionPredominant host speciesGenotypeReference strainReference accession
Hepeviridae BurmaM73218Human
Orthohepevirus BurmaM73218Human
Orthohepevirus ABurmaM73218HumanHEV-1BurmaM73218
HumanHEV-2MexicoM74506
Human, pig, rabbit, deer, mongooseHEV-3MengM74506
Human, pigHEV-4T1AJ272108
Wild boarHEV-5JBOAR135-Shiz09AB573435
Wild boarHEV-6wbJOY_06AB602441
CamelHEV-7DcHEV-178CKJ496143
Orthohepevirus BF93-5077AY535004Chicken
Orthohepevirus CR63GU345042RatHEV-C1R63GU345042
FerretHEV-C2FRHEV4JN998606
Orthohepevirus DBatHEV/BS7/GE/2009JQ001749Bat
Piscihepevirus Heenan LakeHQ731075Trout
Piscihepevirus AHeenan LakeHQ731075Trout

We propose that the remaining HEV variants should be considered as members of the genus Orthohepevirus (Table 1), extending the usage of a previous proposal (Meng, 2013). This change of genus name from Hepevirus is preferred because it makes it clear that not all members of the Hepeviridae belong to the same genus.

Analysis of part of the RdRP domain from an incomplete virus genome sequence (Hepelivirus) obtained from raw sewage suggests that this may represent an additional genus within the family Hepeviridae (Ng et al., 2012). However, a complete genome sequence would be required to confirm the taxonomic position of this virus, for which the host range remains unknown.

Species

The next obvious level of sequence diversity amongst members of the genus Orthohepevirus is that which separates variants originally isolated from different host species (Fig. 3). For example, all isolates from chickens belong to one clade, as do those from rats and ferrets, those from humans, deer, mongooses, pigs, wild boar and camels, while a single complete genome sequence isolated from a bat forms a fourth group. These four groupings are observed regardless, whether the analysis is performed on complete genome nucleotide sequences or subgenomic amino acid sequences (Figs 1 and 3). We propose that these groupings correspond to four species and that they be designated Orthohepevirus A to Orthohepevirus D (Table 1).

Phylogenetic analyses of members of the genus Orthohepevirus. Maximum-likelihood tree for amino acid sequences in (a) ORF1 and (b) ORF2, (c) ORF1-779 to the end of ORF2 with the addition of KF951328 from moose, (d) ORF1-1420 to ORF1-1505 with the addition of sequences from mink, fox, greater bandicoot, Asian musk shrew and bat. Maximum-likelihood trees were computed using the model according to Le & Gascuel with a gamma distribution of evolutionary rates among sites, with some invariant sites. Branches supported by >70 % of bootstrap replicates are indicated.

Subgenomic sequences of variants isolated from other mammalian species have recently become available but phylogenetic analyses are complicated because they overlap with only relatively short regions. The most complete sequence (5081 nt) is derived from a variant isolated from a moose (Lin et al., 2014). Phylogenetic analysis of the amino acid sequence of this isolate for part of ORF1 and all of ORF2 (Fig. 3) shows that it is distinct from all other variants, although it shares a branch with species Orthohepevirus A in the same way that the ferret variants group with rat-derived species Orthohepevirus C. Similar analysis of virus sequences isolated from mink (Krog et al., 2013) shows that these group with sequences isolated from ferrets. A sequence isolated from a fox (Bodewes et al., 2013) appears to be more distinct (Fig. 3d) and could represent an additional species. However, this last analysis was based on a relatively short subgenomic region (ORF1-1420 to ORF1-1505) in which the grouping of species Orthohepevirus C isolates from rat and ferret, and of the moose variant with Orthohepevirus A variants from human, pig and wild boar, were less distinct than for comparisons based on longer coding regions (Fig. 3). We recommend that assignment of the moose and fox variants to particular Orthohepevirus species should not be made until comparisons based upon complete genome sequences are available.

An advantage of using species names of the form Orthohepevirus A, etc., rather than names based on avian HEV, rat HEV or bat HEV, is that they will not be compromised by current or future discoveries on the extent of host range. For example, variants of species Orthohepevirus C have been isolated from Rattus (Li et al., 2013) and Bandicota species (the greater bandicoot rat) (Li et al., 2013), which are members of the order Rodentia, but also from Asian musk shrews (Guan et al., 2013), which are members of the order Soricomorpha, and from ferrets (Li et al., 2014; Raj et al., 2012) and mink (Krog et al., 2013), which are members of the order Carnivora. Wider screening for members of this species, and also for members of species Orthohepevirus B and Orthohepevirus D, may reveal wider host ranges than those recognized at present.

Genotypes

Although the International Committee on Taxonomy of Viruses does not provide official designations for taxonomic entities below the species level, it is useful for researchers to have an agreed designation of genotypes within each species grouping.

<italic>Orthohepevirus</italic> A.

Within species Orthohepevirus A, four genotypes are currently described that infect humans (HEV-1, HEV-2, HEV-3 and HEV-4), and assignment of complete genome sequences to these genotypes is generally unambiguous. The only exceptions are recombinant viruses, which have been documented both within and between genotypes (Chen et al., 2012; van Cuyck et al., 2005; Fan, 2009; Wang et al., 2010), although in some cases recombinant viruses may represent laboratory artefacts (Wang et al., 2010). More problematic is the designation of additional genotypes HEV-5 and HEV-6, names which have been variously assigned to avian HEV and rat HEV (here proposed to be classified as members of the species Orthohepevirus B and Orthohepevirus C, respectively) and variants isolated from wild boar (Oliveira-Filho et al., 2013; Smith et al., 2013; Takahashi et al., 2011), or by implication to variants isolated from rabbits (Geng et al., 2011; Zhao et al., 2009).

Phylogenetic analysis of the nucleotide and amino acid sequences of concatenated ORF1 and ORF2 (excluding the HVR and the hypervariable insertion found in all rabbit HEV isolates) for these variants, including a divergent variant recently described from wild boar (Takahashi et al., 2014), reveals four fully supported branches consisting of (i) HEV-1 and HEV-2, (ii) HEV-3 and variants isolated from rabbits, together with a closely related patient sequence, and (iii) HEV-4 together with all three isolates from wild boar and (iv) isolates from camels (Fig. 4). However, given the precedence of HEV-1 and HEV-2 forming well recognized and phylogenetically distinct genotypes, the least disruptive way of representing these phylogenetic relationships would be to retain the HEV-1, HEV-2, HEV-3 and HEV-4 genotype assignments.

Phylogenetic analyses of members of Orthohepevirus A. Maximum-likelihood trees were produced for 137 concatenated ORF1 and ORF2 regions (excluding the HVR and rabbit HEV-specific insertion) for (a) nucleotide sequences using the general time-reversible model with gamma distribution and invariant sites and (b) amino acid sequences using the Jones–Taylor–Thornton model with frequencies, and gamma distribution with invariant sites. Branches supported by >70 % of bootstrap replicates are indicated.

Variants derived from rabbits (and the single related human isolate JQ013793) show a wider range of amino acid sequence distances from each other (maximum value 0.081) than HEV-1, -3 and -4 (maximum values 0.041, 0.053 and 0.053 respectively) while minimum distances between them and HEV-3 variants are lower (0.061) than between rabbit sequences and other HEV genotypes or variants (minimum value 0.108). A previous study discussed whether rabbit viruses might be assigned to genotype 3 as divergent members or form a separate genotype, since nucleotide and amino acid distances were intermediate between those observed within HEV-1, HEV-3 and HEV-4 and distances between genotypes (Smith et al., 2013). Our analysis here of further rabbit-derived HEV variants (JX121233, JQ768461, JX109834, AB740220, AB740221, AB740222 and JX565469) confirms this finding, for example, the pairwise amino acid distances with genotype 3 that overlap those between genotype 3 and 4. However, the most extreme distances between rabbit HEV and HEV-3 involve the isolates FJ906895 and FJ906896. These sequences contain numerous amino acid substitutions clustered at the C-terminus (FJ906895) or N-terminus (FJ906896), which are at sites that are otherwise highly conserved throughout Orthohepevirus A sequences, and so are likely to represent sequencing artefacts. Excluding these two sequences from the analysis of pairwise distances, there was no overlap between amino acid distances between rabbit and HEV-3 sequences, and between HEV-3 and HEV-4 (Fig. 5). Retaining these isolates but excluding the terminal regions containing the aberrant sites from the Orthohepevirus alignment similarly eliminates overlap between these categories (data not shown). Since HEV-3 and the rabbit HEV also share a long branch on phylogenetic analysis, we consider it simplest to provisionally assign the rabbit sequences to genotype HEV-3.

Frequency distribution of distances among sequences of variants in the genus Orthohepevirus. Histograms show the frequency of nucleotide (a) and amino acid sequence p-distances (b) among 135 concatenated ORF1 and ORF2 sequences (excluding the HVR and rabbit HEV-specific HVR insertion). Maximum and minimum intra-genotype ranges of distances within genotypes 1, 3 and 4 and between rabbit sequences are indicated with open bars. Grey-filled bars depict distances between HEV-3 and rabbit variants (R/3) and inter-genotype distances (3/4, HEV-3 and HEV-4; 1/2, HEV-1 and HEV-2; W/4, wild boar and HEV-4). Distances between the three wild boar sequences are indicated with spots.

On this basis, amino acid distances of concatenated ORF1 and ORF2 (lacking hypervariable regions) greater than 0.088 could then act as a threshold to demarcate intra- and inter- genotype distances. Using this criterion, the three wild boar isolates would comprise two additional genotypes HEV-5 (AB573435) and HEV-6 (AB602441 and AB856243 differing from each other by 0.076 and from HEV-5 by >0.10), while the variants isolated from camels (differing from all other sequences by >0.095) would become HEV-7.

The virus isolated from moose appears to constitute a distinct lineage within species Orthohepevirus A, but definitive placement requires a complete genome sequence.

<italic>Orthohepevirus</italic> B.

Although four genotypes of Orthohepevirus B have been proposed (Bilic et al., 2009; Hsu & Tsai, 2014; Huang et al., 2004), these are much less diverse than the genotypes of Orthohepesvirus A. For example there is <6 % divergence in Orthohepevirus B complete genome amino acid sequences (Figs 1 and 3), which is less than the divergence observed within HEV-3 (<9 %) and HEV-4 (<7 %). In addition, the amino acid sequence distances among the eight currently available complete genome sequences of members of species Orthohepevirus B form a continuous distribution, where distances within genotypes (maximum 3.3 %) approach those between genotypes (minimum 3.6 %). If additional complete genome sequences were available, this narrow division might disappear.

<italic>Orthohepevirus</italic> C.

The extent of diversity (<11 %) among complete genome amino acid sequences of the rat-derived Orthohepevirus C variants barely overlaps that observed among genotypes of species Orthohepevirus A (10–18 %). However, much greater divergence is observed between rat and ferret Orthohepevirus C variants (23 %). Analysis of the short region of ORF1 for which sequence information is available from additional variants (Fig. 2d) indicates that diversity among greater bandicoot and Asian musk shrew isolates falls within that of the rat variants (Guan et al., 2013; Li et al., 2013), while isolates from the mink group falls within those from ferret (Krog et al., 2013). Based on this information, we propose that Orthohepevirus C may be divided into two genotypes, namely HEV-C1, including isolates derived from hosts in the orders Rodentia and Soricomorpha, and HEV-C2, including isolates derived from ferret (and possibly mink).

<italic>Orthohepevirus</italic> D.

A single complete genome sequence is available for species Orthohepevirus D, but phylogenetic analysis of a short region of ORF2, for which data from additional isolates are available, suggests a level of diversity equivalent to that within species Orthohepevirus A and Orthohepevirus C (Fig. 2d). While this is consistent with the existence of multiple genotypes within Orthohepevirus D, additional sequence information is required to confirm that these relationships prevail for larger genomic regions.

Subgenotypes

Several studies have attempted to define subgenotypes of genotypes HEV-1, HEV-3 and HEV-4 in the species Orthohepevirus A, in some instances based on the analysis of subgenomic regions (Lu et al., 2006; Zhu et al., 2011). Such categories may form useful labels for epidemiological studies (Dai et al., 2013), but more recent analysis of complete genome sequences suggests that it is not possible to define discrete boundaries that distinguish subgenotypes with consistency (Okamoto, 2007; Oliveira-Filho et al., 2013; Smith et al., 2013). We recommend the approach, commonly adopted in several recent publications, of labelling clades apparent within sequence sets (Dai et al., 2013; Ijaz et al., 2014; Oliveira-Filho et al., 2013) without defining them as permanent classification assignments.

Reference sequences and numbering

A recurring difficulty in the literature, relating to molecular studies of members of the family Hepeviridae, is that of comparing different studies for which there is no explicit standard sequence with reference to which nucleotides or amino acid residues are numbered. The presence of numerous insertions or deletions and regions of low similarity in alignments of Hepeviridae sequences precludes a unified numbering system that is applicable across all species or genera. We recommend that genome sequences be numbered with reference to the first nucleotide of the prototype complete genome sequence available for each species within the genus Orthohepevirus (Table 1). Nucleotide sites in variants that contain insertions relative to the prototype sequence should be identified with additional letters, beginning at the site of insertion. For example, a three-nucleotide insertion at position 1788 of the prototype sequence would be numbered 1788a, 1788b, 1788c. Insertions of more than 26 nt would be numbered from the twenty-seventh position as 1788aa, 1788ab, etc. and then as 1788ba, 1788bb, etc. as required. This mirrors the system adopted for HCV (Kuiken et al., 2006). Similarly, amino acid residues should be numbered with reference to the first residue of the appropriate ORF from the reference sequence, for example ORF1-929, with an insertion at this site indicated by suffix letters such as ORF1-929a, ORF1-929b, etc., followed if necessary by ORF1-929 aa, ORF1-929ab, etc.

Conclusions

The proposed classification, which assigns a separate hierarchy of genus and species, respects the different levels of divergence between the cutthroat trout virus and all other hepeviruses. The degrees of sequence divergence and conservation of genomic features associated with these categories closely match those attributed to genera and species in other virus families (e.g. Picornaviridae, Caliciviridae and Flaviviridae). This description of relationships among members of the family Hepeviridae will help to resolve current confusion in the literature and, by providing a rational basis for taxonomic assignments, help to reduce the number of future conflicts as more members of this family are discovered.

Methods

Phylogenetic analysis included the complete genome sequences in GenBank accessions M73218, M74506, AF082843, FJ906895, AB301710, AJ272108, AB602441, AB856243, AB573435, KJ496143, KJ496144, JN167537, JN167538, GU345042, GU345043, JX120573, AB847305, AB847306, AB847307, AB847308, AB847309, JN998606, JN998607, AB890001, AB890374, JN597006, JN997392, AM943647, AM943646, GU954430, AY535004, EF206691, KF511797, JQ001749 and HQ731075. Analyses of subgenomic regions included the additional sequences JQ001744-8 and JQ071861 (bat), JN167530-6, KC473527-31, JN040433, KC294199, JF516246, GQ504009-10 and AB725884-900 and AB847310-406 (rat), KF268376-93 (ferret), KC465990-6001 (greater bandicoot and Asian musk rat), KC692369-70 (fox), KC802090-3 (mink), AY043166 (chicken) and KF951328 (moose).

A further dataset included 137 complete genome sequences isolated from humans, pigs, rabbits and wild boar (downloaded from GenBank on 4 February 2014, excluding recombinant sequences and sequences differing from other sequences by less than 0.2 % of nucleotide positions in ORF1), as follows: M73218, AB720034, JF443717, JF443718, JF443720, JF443721, JF443722, JF443723, JF443725, JF443726, JQ655734, FJ457024, D11092, AY204877, AY230202, AF185822, AF076239, X99441, M74506, KC618402, AB780450, AB740232, JQ953664, JQ953665, JQ953666, JN837481, JN906974, AB593690, AB630970, AB630971, AB591733, AB591734, HQ389543, HQ709170, AB481226, AB481228, AB481229, FJ653660, FJ426403, FJ426404, FJ956757, FJ998008, FJ705359, AB291951, AB291953, FJ527832, EU375463, AB189071, AB236320, EU723512, EU723514, AB073912, EU495148, AB089824, AB091394, AB222182, AB222183, AB222184, AB248520, AB248522, AB290312, AB290313, AB291961, AB291962, AB291963, EU723513, EU723516, AB074920, AP003430, EU360977, AB369687, AB369689, AB369691, AB246676, AY575857, AF455784, AY115488, AF060669, KF922359, AF082843, FJ906895, JX109834, AB740220, FJ906896, AB740221, AB740222, JX565469, JQ013791, JQ013792, JQ013793, FJ610232, KC492825, KF176351, JF915746, JQ740781, JX855794, AB291959, JQ655733, JQ655735, JQ655736, AB698654, JQ993308, AB602439, GU361892, GU119960, GU119961, GU206559, HQ634346, AB481227, HM439284, FJ763142, GU188851, AB480825, AB197673, AB197674, AB091395, AB220974, AB291964, AB074915, AB080575, EU676172, AB108537, AB369688, AB369690, EU366959, EF570133, DQ279091, EF077630, AY723745, AB253420, AY594199, AJ272108, AB602441, AB856243, AB573435, KJ496143 and KJ496144. Sequences were aligned using muscle v3.8 (Edgar, 2004) within sse v1.1 (Simmonds, 2012), and then refined manually. Phylogenetic analysis was conducted using mega version 6 (Tamura et al., 2013). Using the optimal model for each dataset, maximum-likelihood trees were reproduced by using the programs Models and Phylogeny in mega 6. Distances between nucleotide and amino acid sequences were generated within sse program. Homology to known protein domains were identified using Motif Scan (http://myhits.isb-sib.ch).

Acknowledgements

We are grateful to Dr Patrick Woo for sharing the camel HEV isolate complete genome sequences ahead of general release. This research was supported by a grant from The Wellcome Trust to the Centre for Immunity, Infection and Evolution at the University of Edinburgh. This information is distributed solely for the purpose of pre-dissemination peer review under applicable information quality guidelines. It has not been formally disseminated by the Centers for Disease Control and Prevention/Agency for Toxic Substances and Disease Registry. It does not represent and should not be construed to represent any agency determination or policy. Use of trade names is for identification only and does not imply endorsement by the US Department of Health and Human Services, the Public Health Service, or the Centers for Disease Control and Prevention. The findings and conclusions in this report are those of the authors and do not necessarily represent the views of the Centers for Disease Control and Prevention. The authors declare that they have no conflict of interest.

ReferencesBattsW.YunS.HedrickR.WintonJ. (2011). A novel member of the family Hepeviridae from cutthroat trout (Oncorhynchus clarkii). Virus Res 158, 116123 10.1016/j.virusres.2011.03.01921458509BilicI.JaskulskaB.BasicA.MorrowC. J.HessM. (2009). Sequence analysis and comparison of avian hepatitis E viruses from Australia and Europe indicate the existence of different genotypes. J Gen Virol 90, 863873 10.1099/vir.0.007179-019264623BodewesR.van der GiessenJ.HaagmansB. L.OsterhausA. D. M. E.SmitsS. L. (2013). Identification of multiple novel viruses, including a parvovirus and a hepevirus, in feces of red foxes. J Virol 87, 77587764 10.1128/JVI.00568-1323616657ChenX.ZhangQ.HeC.ZhangL.LiJ.ZhangW.CaoW.LvY.-G.LiuZ. & other authors (2012). Recombination and natural selection in hepatitis E virus genotypes. J Med Virol 84, 13961407 10.1002/jmv.2323722825818DaiX.DongC.ZhouZ.LiangJ.DongM.YangY.FuJ.TianH.WangS. & other authors (2013). Hepatitis E virus genotype 4, Nanjing, China, 2001-2011. Emerg Infect Dis 19, 15281530 10.3201/eid1909.13001323965731DrexlerJ. F.SeelenA.CormanV. M.Fumie TatenoA.CottontailV.Melim ZerbinatiR.Gloza-RauschF.KloseS. M.Adu-SarkodieY. & other authors (2012). Bats worldwide carry hepatitis E virus-related viruses that form a putative novel genus within the family Hepeviridae. J Virol 86, 91349147 10.1128/JVI.00800-1222696648EdgarR. C. (2004). MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32, 17921797 10.1093/nar/gkh34015034147FanJ. (2009). Open reading frame structure analysis as a novel genotyping tool for hepatitis E virus and the subsequent discovery of an inter-genotype recombinant. J Gen Virol 90, 13531358 10.1099/vir.0.009431-019264661GengJ.FuH.WangL.BuQ.LiuP.WangM.SuiY.WangX.ZhuY.ZhuangH. (2011). Phylogenetic analysis of the full genome of rabbit hepatitis E virus (rbHEV) and molecular biologic study on the possibility of cross species transmission of rbHEV. Infect Genet Evol 11, 20202025 10.1016/j.meegid.2011.09.00621945590GuanD.LiW.SuJ.FangL.TakedaN.WakitaT.LiT.-C.KeC. (2013). Asian musk shrew as a reservoir of rat hepatitis E virus, China. Emerg Infect Dis 19, 13411343 10.3201/eid1908.13006923876489HsuI. W.-Y.TsaiH.-J. (2014). Avian hepatitis E virus in chickens, Taiwan, 2013. Emerg Infect Dis 20, 149151 10.3201/eid2001.13122424378180HuangC.-C.NguyenD.FernandezJ.YunK. Y.FryK. E.BradleyD. W.TamA. W.ReyesG. R. (1992). Molecular cloning and sequencing of the Mexico isolate of hepatitis E virus (HEV). Virology 191, 550558 10.1016/0042-6822(92)90230-M1448913HuangF. F.SunZ. F.EmersonS. U.PurcellR. H.ShivaprasadH. L.PiersonF. W.TothT. E.MengX. J. (2004). Determination and analysis of the complete genomic sequence of avian hepatitis E virus (avian HEV) and attempts to infect rhesus monkeys with avian HEV. J Gen Virol 85, 16091618 10.1099/vir.0.79841-015166445IjazS.SaidB.BoxallE.SmitE.MorganD.TedderR. S. (2014). Indigenous hepatitis E in England and Wales from 2003 to 2012: evidence of an emerging novel phylotype of viruses. J Infect Dis 209, 12121218 10.1093/infdis/jit65224273173JohneR.HeckelG.Plenge-BönigA.KindlerE.MareschC.ReetzJ.SchielkeA.UlrichR. G. (2010a). Novel hepatitis E virus genotype in Norway rats, Germany. Emerg Infect Dis 16, 14521455 10.3201/eid1609.10044420735931JohneR.Plenge-BönigA.HessM.UlrichR. G.ReetzJ.SchielkeA. (2010b). Detection of a novel hepatitis E-like virus in faeces of wild rats using a nested broad-spectrum RT-PCR. J Gen Virol 91, 750758 10.1099/vir.0.016584-019889929KrogJ. S.BreumS. Ø.JensenT. H.LarsenL. E. (2013). Hepatitis E virus variant in farmed mink, Denmark. Emerg Infect Dis 19, 20282030 10.3201/eid1912.13061424274600KuikenC.CombetC.BukhJ.Shin-IT.DeleageG.MizokamiM.RichardsonR.SablonE.YusimK. & other authors (2006). A comprehensive system for consistent numbering of HCV sequences, proteins and epitopes. Hepatology 44, 13551361 10.1002/hep.2137717058236LackJ. B.VolkK.Van Den BusscheR. A. (2012). Hepatitis E virus genotype 3 in wild rats, United States. Emerg Infect Dis 18, 12681273 10.3201/eid1808.12007022840202LiW.GuanD.SuJ.TakedaN.WakitaT.LiT.-C.KeC. W. (2013). High prevalence of rat hepatitis E virus in wild rats in China. Vet Microbiol 165, 275280 10.1016/j.vetmic.2013.03.01723623690LiT.-C.YangT.AmiY.SuzakiY.ShirakuraM.KishidaN.AsanumaH.TakedaN.TakajiW. (2014). Complete genome of hepatitis E virus from laboratory ferrets. Emerg Infect Dis 20, 709712 10.3201/eid2004.13181524655541LinJ.NorderH.UhlhornH.BelákS.WidénF. (2014). Novel hepatitis E like virus found in Swedish moose. J Gen Virol 95, 557570 10.1099/vir.0.059238-024296469LuL.LiC.HagedornC. H. (2006). Phylogenetic analysis of global hepatitis E virus sequences: genetic diversity, subtypes and zoonosis. Rev Med Virol 16, 536 10.1002/rmv.48216175650MengX.-J. (2013). Zoonotic and foodborne transmission of hepatitis E virus. Semin Liver Dis 33, 4149 10.1055/s-0033-133811323564388NgT. F. F.MarineR.WangC.SimmondsP.KapusinszkyB.BodhidattaL.OderindeB. S.WommackK. E.DelwartE. (2012). High variety of known and new RNA and DNA viruses of diverse origins in untreated sewage. J Virol 86, 1216112175 10.1128/JVI.00869-1222933275OkamotoH. (2007). Genetic variability and evolution of hepatitis E virus. Virus Res 127, 216228 10.1016/j.virusres.2007.02.00217363102Oliveira-FilhoE. F.KönigM.ThielH.-J. (2013). Genetic variability of HEV isolates: inconsistencies of current classification. Vet Microbiol 165, 148154 10.1016/j.vetmic.2013.01.02623434129RajV. S.SmitsS. L.PasS. D.ProvaciaL. B. V.Moorman-RoestH.OsterhausA. D. M. E.HaagmansB. L. (2012). Novel hepatitis E virus in ferrets, The Netherlands. Emerg Infect Dis 18, 13691370 10.3201/eid1808.11165922840220ReyesG. R.PurdyM. A.KimJ. P.LukK. C.YoungL. M.FryK. E.BradleyD. W. (1990). Isolation of a cDNA from the virus responsible for enterically transmitted non-A, non-B hepatitis. Science 247, 13351339 10.1126/science.21075742107574SimmondsP. (2012). SSE: a nucleotide and amino acid sequence analysis platform. BMC Res Notes 5, 50 10.1186/1756-0500-5-5022264264SimmondsP.BukhJ.CombetC.DeléageG.EnomotoN.FeinstoneS.HalfonP.InchauspéG.KuikenC. & other authors (2005). Consensus proposals for a unified system of nomenclature of hepatitis C virus genotypes. Hepatology 42, 962973 10.1002/hep.2081916149085SmithD. B.PurdyM. A.SimmondsP. (2013). Genetic variability and the classification of hepatitis E virus. J Virol 87, 41614169 10.1128/JVI.02762-1223388713SmithD. B.BukhJ.KuikenC.MuerhoffA. S.RiceC. M.StapletonJ. T.SimmondsP. (2014). Expanded classification of hepatitis C virus into 7 genotypes and 67 subtypes: updated criteria and genotype assignment web resource. Hepatology 59, 318327 10.1002/hep.2674424115039TakahashiM.NishizawaT.SatoH.SatoY.JirintaiNagashimaS.OkamotoH. (2011). Analysis of the full-length genome of a hepatitis E virus isolate obtained from a wild boar in Japan that is classifiable into a novel genotype. J Gen Virol 92, 902908 10.1099/vir.0.029470-021228128TakahashiM.NishizawaT.NagashimaS.JirintaiS.KawakamiM.SonodaY.SuzukiT.YamamotoS.ShigemotoK. & other authors (2014). Molecular characterization of a novel hepatitis E virus (HEV) strain obtained from a wild boar in Japan that is highly divergent from the previously recognized HEV strains. Virus Res 180, 5969 10.1016/j.virusres.2013.12.01424370869TamA. W.SmithM. M.GuerraM. E.HuangC.-C.BradleyD. W.FryK. E.ReyesG. R. (1991). Hepatitis E virus (HEV): molecular cloning and sequencing of the full-length viral genome. Virology 185, 120131 10.1016/0042-6822(91)90760-91926770TamuraK.StecherG.PetersonD.FilipskiA.KumarS. (2013). MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evol 30, 27252729 10.1093/molbev/mst19724132122TsarevS. A.EmersonS. U.ReyesG. R.TsarevaT. S.LegtersL. J.MalikI. A.IqbalM.PurcellR. H. (1992). Characterization of a prototype strain of hepatitis E virus. Proc Natl Acad Sci U S A 89, 559563 10.1073/pnas.89.2.5591731327van CuyckH.FanJ.RobertsonD. L.RoquesP. (2005). Evidence of recombination between divergent hepatitis E viruses. J Virol 79, 93069314 10.1128/JVI.79.14.9306-9314.200515994825WangH.ZhangW.NiB.ShenH.SongY.WangX.ShaoS.HuaX.CuiL. (2010). Recombination analysis reveals a double recombination event in hepatitis E virus. Virol J 7, 129 10.1186/1743-422X-7-12920546601WooP. C.LauS. K.TengJ. L.TsangA. K.JosephM.WongE. Y.TangY.SivakumarS.XieJ. & other authors (2014). New hepatitis E virus genotype in camels, the Middle East. Emerg Infect Dis 20, 10441048 10.3201/eid2006.14014024856611ZhaoC.MaZ.HarrisonT. J.FengR.ZhangC.QiaoZ.FanJ.MaH.LiM. & other authors (2009). A novel genotype of hepatitis E virus prevalent among farmed rabbits in China. J Med Virol 81, 13711379 10.1002/jmv.2153619551838ZhuY. M.DongS. J.SiF. S.YuR. S.LiZ.YuX. M.ZouS. X. (2011). Swine and human hepatitis E virus (HEV) infection in China. J Clin Virol 52, 155157 10.1016/j.jcv.2011.06.02321831705