key: cord-0019457-hktk14tg authors: Tohma, Kentaro; Lepore, Cara J.; Martinez, Magaly; Degiuseppe, Juan I.; Khamrin, Pattara; Saito, Mayuko; Mayta, Holger; Nwaba, Amy U. Amanda; Ford-Siltz, Lauren A.; Green, Kim Y.; Galeano, Maria E.; Zimic, Mirko; Stupka, Juan A.; Gilman, Robert H.; Maneekarn, Niwat; Ushijima, Hiroshi; Parra, Gabriel I. title: Genome-wide analyses of human noroviruses provide insights on evolutionary dynamics and evidence of coexisting viral populations evolving under recombination constraints date: 2021-07-13 journal: PLoS Pathog DOI: 10.1371/journal.ppat.1009744 sha: 8b7d70159f20a9cdf79aa0dbb1213c7b4bb1d50f doc_id: 19457 cord_uid: hktk14tg Norovirus is a major cause of acute gastroenteritis worldwide. Over 30 different genotypes, mostly from genogroup I (GI) and II (GII), have been shown to infect humans. Despite three decades of genome sequencing, our understanding of the role of genomic diversification across continents and time is incomplete. To close the spatiotemporal gap of genomic information of human noroviruses, we conducted a large-scale genome-wide analyses that included the nearly full-length sequencing of 281 archival viruses circulating since the 1970s in over 10 countries from four continents, with a major emphasis on norovirus genotypes that are currently underrepresented in public genome databases. We provided new genome information for 24 distinct genotypes, including the oldest genome information from 12 norovirus genotypes. Analyses of this new genomic information, together with those publicly available, showed that (i) noroviruses evolve at similar rates across genomic regions and genotypes; (ii) emerging viruses evolved from transiently-circulating intermediate viruses; (iii) diversifying selection on the VP1 protein was recorded in genotypes with multiple variants; (iv) non-structural proteins showed a similar branching on their phylogenetic trees; and (v) contrary to the current understanding, there are restrictions on the ability to recombine different genomic regions, which results in co-circulating populations of viruses evolving independently in human communities. This study provides a comprehensive genetic analysis of diverse norovirus genotypes and the role of non-structural proteins on viral diversification, shedding new light on the mechanisms of norovirus evolution and transmission. a1111111111 a1111111111 a1111111111 a1111111111 a1111111111 Norovirus is the major cause of acute gastroenteritis in all age groups. The most common symptoms of norovirus infection are diarrhea, vomiting, nausea, and abdominal cramps. In healthy individuals the disease is resolved rapidly, but in vulnerable populations (e.g. the elderly, malnourished children, or immunocompromised patients) symptoms can last longer, resulting in severe dehydration and death. Annually, norovirus is associated with over 600 million cases of acute gastroenteritis and up to 200,000 deaths in children, mostly from developing countries [1, 2] . Thus, a vaccine could save hundreds of thousands of lives and help to mitigate the burden of norovirus disease. The human norovirus genome is a single-stranded, positive-sense, polyadenylated, RNA molecule of~7.5kb in size, organized into three open reading frames (ORFs) flanked by two short untranslated regions. ORF1 encodes the non-structural proteins (NS1/2-7), required for virus replication, while ORF2 and ORF3 encode the major (VP1) and minor (VP2) capsid proteins, respectively [3] . The genome is enclosed by an array of capsid proteins that form icosahedral particles with a diameter of approximately 27 nm [4] . The structural model of norovirus VP1 presents two domains, the shell (S) and the protruding (P) domain [5] . The S domain forms the scaffold of the icosahedral capsid, while the P domain is a projection that contains the determinants of host interactions, including neutralizing epitopes and carbohydrate binding sites that act as attachment factors to facilitate infection [6] [7] [8] . While the precise role of VP2 remains to be elucidated, there is evidence that it might be involved in particle stabilization [9] . One of the major obstacles for norovirus vaccine development is the large viral genetic and antigenic diversity [10] . Genetic differences within VP1 have been used to classify noroviruses into genogroups and genotypes [11] . Over 30 different genotypes, mostly from genogroups I (GI) and II (GII), have been shown to infect humans. While most genotypes circulate with variable incidences, a single genotype (GII.4) predominates globally [12, 13] . This dominance of GII.4 viruses has been explained by the accrual of mutations on the VP1, resulting in the emergence of new variants [14] [15] [16] . Although GII.4 predominance has been recorded for over two decades, other genotypes can transiently predominate in a given geographical location. This is evidenced by the recent emergence and increase in incidence of GII.17 and GII.2 in different countries [17] [18] [19] [20] [21] . While these two viruses are antigenically distinct to GII.4 [8, 22] , suggesting immune escape as a mechanism of emergence, it seems that norovirus emergence is a multifactorial event that includes changes in viral proteins other than VP1 and virus-host interactions [12] . Thus, the emergence of epidemic GII.2 viruses was associated with changes in the NS7 (viral RNA polymerase) [23] , and the emergence of epidemic GII.17 viruses was associated with mutations in VP1 that resulted in the emergence of a new variant and changes in host susceptibility [24] [25] [26] . It has been shown that noroviruses are prone to recombine at the ORF1/ORF2 boundaries [27] . Thus, because norovirus also presents different types of viral RNA polymerases [11] , viruses with different combinations of capsid and polymerase types have been detected in nature [27] [28] [29] . While most recombinant viruses have been found circulating at low levels in the human population, emergence of epidemic viruses has also been associated with recombination events. The most recent example is the emergence of different viruses associated with the GII.P16 polymerase type causing outbreaks around the world [30] . It seems that mutations in this viral polymerase provided new characteristics for successful human infection, as shown by the large outbreaks of GII.2[P16] (GII.2 capsid and GII.P16 polymerase) viruses reported in the 2016/2017 season. This polymerase was initially associated with GII.2 and GII.4 capsids [17, 31] , but recently this polymerase has been reported with multiple other capsids, including GII.1, GII.3, GII.12, and GII. 13 [30,32] . The mechanism by which the viral polymerase is involved in viral emergence is not completely understood, but might include differences in the mutational and/or transcriptional rate that could facilitate transmission at the individual or population level [33] [34] [35] . Several studies have utilized full-genome sequences to evaluate norovirus transmission and evolution in different settings. Thus, full-genome sequences of noroviruses have proven to be instrumental in assessing the diversification and the direction of transmission during nosocomial infections [36] [37] [38] [39] [40] and given geographical regions [41] [42] [43] . Importantly, some of these studies have shown that, in addition to changes in the major structural protein, VP1, and the viral RNA polymerase, other viral proteins (like NS1/2 or VP2) might play an important role in the transmission and emergence of novel noroviruses [30, 42] . Despite almost two decades of robust sequence and epidemiological norovirus studies, such full-length genome analyses are scarce and genomic regions other than VP1 and NS7 have been largely understudied. Moreover, most norovirus genome records available were obtained from predominant genotypes that circulated since the mid-2000s in developed countries. Thus, there is a large data gap on the spatiotemporal genomic diversification of human noroviruses. To gain insights on the evolution and mechanisms of norovirus emergence, we collected archival samples and sequenced over 281 viruses circulating since the 1970s in over 10 countries from four continents, with a major emphasis on those genotypes that are underrepresented in the public databases. Thus, using this improved dataset of human norovirus genomes we aimed to (i) revisit evolutionary parameters at the genome-wide level that could provide insights on norovirus diversification, (ii) perform a detailed analysis of the phylogenetic relationship among atypical and untypeable viruses, and (iii) investigate factors that govern recombination among noroviruses. 1972 to 2019 (Fig 1 and S1 Table) . The viruses sequenced here were mostly sampled from geographic regions and decades with scarce information. Our current sequencing efforts include an expansion of the geographical information, with a particular increase in the number of norovirus genomes for South America (27 to 169 genomes), Africa (14 to 22 genomes), and Asia (1,082 to 1,169 genomes), the first two geographic regions with very under-represented norovirus genomic information (Fig 1A) . Our efforts also increased the number of genomes for noroviruses sampled before 2000 (62 to 144 genomes) (Fig 1B) , providing the oldest genome information from 12 norovirus genotypes (Table 1) , and new genomes for genotypes with under-represented genomic information (Fig 1C) . Thus, this study provides more than 42% (135/325) of the genome information available for those under-represented human norovirus genotypes (Fig 1C) , with new nearly full-length genomes for genotypes GII.9, GII. 27 [44] , and GII.NA2 [45] . Together with the genomes retrieved from GenBank, our analyses included a dataset of 2,013 human norovirus genomes. Due to the evolutionary patterns presented by the different norovirus genotypes [14, 46] , absence of viruses circulating in different decades could result in inconsistencies on the estimates of evolutionary rates [21] , particularly for those genotypes presenting different variants [46] . Thus, because our work provided the oldest genome information for 12 out of the 36 genotypes described for human norovirus, we thought to calculate the evolutionary rates using this novel information. We found that inclusion of historical samples collected before the 1990s significantly improved the clock-likeness of the VP1-encoding sequences in the dataset (P<0.01 in Wilcoxon matched-pairs signed rank test, Fig 2A) . The R 2 from VP1 phylogenetic root-to-tip regression was improved in correlation with longer duration of years analyzed, providing the accuracy of evolutionary rates estimated. Overall, the rate of evolution was similar for most of the human noroviruses (1.12-4.86 × 10 −3 nucleotide substitutions/site/year), with the 95% highest posterior density (95%HPD) intervals overlapping among most of the genotypes/variants (Fig 2B and S2 Table) . Some of the 95%HPD intervals did not overlap with others, suggesting small differences among the rate of evolution of distinct genotypes. The highest evolutionary rate on VP1-encoding sequences was recorded in GII.4 viruses, which presented periodic emergence and replacement of variants [12] . Some non-GII.4 viruses, e.g. GI.3, GII.6, and GII.17, could be also divided into variants using a cutoff of 5% amino acid differences in the VP1. Notably, while GII.4 variants circulated only for a short period (3-8 years), most non-GII.4 variants have been shown to co-circulate for decades [14, 46] . The distinct non-GII.4 variants presented only minor differences in the rate of evolution, but that difference did not correlate with the prevalence and epidemic potential of viruses. Diversifying selection on the VP1 protein was recorded in major genotypes, with GII.4 viruses presenting the largest number of sites under positive pressure (Fig 3A) . Most of these residues were mapped on the P2 sub-domain, as previously shown [16] . Non-GII.4 genotypes, i.e. GI.3, GII.3, GII.6, GII.17, and GII.21, also presented a large number of sites identified as being under positive pressure, which seems correlated with the number of variants (R 2 = 0.66 in a linear regression) rather than the number of sequences (R 2 = 0.30) or time span available (R 2 = 0.06) (Fig 3B) . Accordingly, the estimated number of sites under positive pressure decreased sharply when looking into individual variants, with a smaller number of sites on P2 domains identified to be under pressure (S1 Fig). Notably, despite the large number of sites were identified as under positive pressure on those non-GII.4 genotypes, none of them accumulated amino acid differences on VP1 as compared to GII.4 noroviruses (Fig 4) [14] . nearly full-length genome sequences highlighted in dark gray. The pie chart indicates the ratio of newly obtained sequences in this study (red) and those from the public database, GenBank (gray). The size of the pie chart shows relative sample size for each country. Geographic map (1:10m Cultural Vector, Admin 0 -Countries) was obtained at https://www.naturalearthdata.com/downloads/10mcultural-vectors/10m-admin-0-countries/ (accessed on June 25, 2021). (B) The line indicates the number of nearly full-length genome Genome-wide analyses of GI and GII viruses revealed that in addition to VP1, NS1/2, NS4, and VP2 proteins presented significantly higher diversity at both nucleotide and amino acid level (P<0.05 in one-way ANOVA with post hoc Tukey multiple-comparison test, Fig 5) . The sequences by year and the bar graph indicates the ratio of newly obtained sequences (red) and public database (gray). (C) Number of newly obtained sequences for major and minor genotypes. The pie chart indicates the ratio of newly obtained sequences in this study (red) and those from public database, GenBank (gray). amino acid differences have been accumulating over time in non-structural proteins, such as NS4 and NS7, and VP2 proteins with the high/low regression slope reflecting its overall diversity (S2- S4 Figs) . Notably, there were no significant differences in the diversification pattern of those proteins between GII.4 and non-GII.4 viruses. The substitution rate of nucleotide sequences from non-structural proteins and VP2-encoding sequences showed an overall similar rate among genome regions and polymerase types/genotypes (Fig 6) , which was within the same range as those from VP1 sequences. As the VP1 protein is the major target of immune responses, changes on this protein could be a key driver of viral transmission [16, 47] . However, while some noroviruses emerge and transiently predominate globally (epidemic), other viruses with the same polymerase and/or capsid types only circulated at low levels (endemic) or were associated with geographically-and temporally-restricted outbreaks. Fig). The phylogenetic trees along with ancestral sequence reconstruction indicated multiple amino acid mutations on the NS4 and NS7 that differentiated the endemic and epidemic viruses on the trees. Other non-structural proteins did not possess specific residues that could separate epidemic and endemic viruses on the trees and those viruses were distributed on the same branches. An exception to searched for mutations on the phylogenetic trees from GII.P4 non-structural proteins that differentiate these variants from those considered epidemic variants; however, neither phylogenetic signals nor amino acid mutations at the non-structural proteins were able to differentiate the epidemic from endemic GII.P4 viruses (S6 Fig) . Interestingly, the phylogenetic diversification of VP2 proteins revealed differences among epidemic and endemic GII. [51] presented a VP2 protein that is relatively close to the Grimsby 1995 variant and a GII.P31 polymerase that clustered with non-epidemic viruses. The epidemiological relevance of the Hong Kong 2019 variant, and associated genes, is currently unclear but will be revealed in the future. Together, we found multiple mutations on NS1/2, NS4, and NS7 proteins that could differentiate the epidemic and endemic viruses. However, none of those mutations were detected in non-structural proteins from epidemic GII.4[P4] viruses, indicating that these different residues could have emerged due to founder effect and/or not by contributing to epidemic potential [12] . In addition to contributing archival full-length genome information of major and minor human norovirus genotypes, our study also focused on expanding the information for atypical viruses, as those could provide valuable information to our understanding of the epidemiology and/or evolution of human noroviruses. Thus, we sequenced a GII.17 virus (Arg13099) detected in 2015 in Argentina that did not cluster with any of the variants described for the VP1 [52] (Fig 7A) . We also found in our dataset two archival GII.17 viruses (T055/Tunisia and DS284/Saudi Arabia from 1977 and 1990, respectively) that were indeed intermediate, and did not cluster with the rest of the viruses of each given variant on the VP1 tree. To gain insights on the diversification of GII.17 viruses, we also analyzed the NS7 sequences from those viruses. The Arg13099 virus did not cluster with any polymerase type, and the T055 virus was typed as GII.P13, but situated outside of the cluster that included other GII.P13 viruses ( Fig 7A) . DS284 was the only GII.17 virus that presented a GII.P3 polymerase. Because GII.P3, GII.P13, and GII.P17 NS7 sequences share a common ancestor and are all genetically close, it is difficult to determine whether the DS284 virus is a recombinant or another lineage. Other intermediate viruses from GII.17 were also reported in the public database, e.g. 27-3/Japan, NORO_231_20/ United Kingdom. Thus, these different intermediate viruses suggest that the GII.17 virus explored multiple variants (and phenotypes) until reaching epidemic potential in 2014. Because of the very low incidence in the human population, GIX.1 (formerly GII.15) viruses are regarded as a minor genotype. This genotype has not been reported to cause large outbreaks and is only detected sporadically in limited studies [53-55]. However, when analyzing samples from US troops deployed to Saudi Arabia in 1990, Desert Shield operations, that Large-scale genome analyses of human noroviruses presented gastrointestinal symptoms [56], we detected 14 GIX.1 viruses in 22 of the samples successfully sequenced from this outbreak ( Fig 7B) . Two distinct transmission chains of GIX.1 viruses were observed in this outbreak, which presented 37-39 nucleotide differences on the VP1 among the viruses from those two transmissions; only two of them were nonsynonymous mutations. Overall, the GIX.1 genotype presented very limited diversity on its VP1 and the corresponding GII.P15 polymerase for the past 30 years (Figs 4 and S3 ). The rest of the samples from the Desert Shield outbreak were characterized as the prototype GI.3 virus (DSV395; Desert Shield virus), GI.7, GII.3, GII.14, and GII.17. This finding suggests that viruses belonging to the GIX.1 genotype could also be linked to large gastrointestinal outbreaks [57] . During our study, we also found novel viruses (recently assigned as GII.26, GII.27, and GII. NA2) [44, 45] . Notably, these viruses belonging to different genotypes clustered together and most of them were detected in different countries from South America [44]. One of the viruses was a recombinant with a VP1 sequence from GII.12, and all of them presented similar VP2 sequences with those from GII. Finally, multiple untypeable viruses were detected when their NS7 were analyzed ( Fig 7C) . While the NS7 protein showed less variation as compared with VP1 sequences (Fig 5) , the untypeable viruses were placed on the tree as intermediates of the evolution among polymerase types. In norovirus nomenclature, the different genotypes are assigned by a tight phylogenetic clustering of viruses, and recombinant viruses are mostly recognized by presenting a VP1 capsid genotype and a different polymerase type. Thus, these intermediate viruses make the polymerase clustering less distinct, and the typing of novel strains more challenging. Recombination is one of the most important aspects of the molecular epidemiology of noroviruses [27] . A large number of recombinant viruses has been reported in the literature [28, 29] , but only a few are regarded as epidemiologically relevant. Some of those are GII. 4 Based on the number of genotypes and polymerase types, recombination at the ORF1 and ORF2 junction region could theoretically generate >1200 capsid-polymerase combinations; however, in our nearly full-length database we only found 92 capsid-polymerase combinations, including 18 new combinations that were confirmed with full-length genome sequences in this study (Table 2) . To gain insights on the mechanisms behind norovirus recombination, we first organized and tabulated all capsid and polymerase types included in our dataset based on their phylogenetic clustering (Fig 8) . Notably, the phylogeny from NS7 presented five major clusters, each including several polymerase types (Fig 8) , that were associated with a specific subset of capsid genotypes. Thus, the GI genotypes presented two distinct clusters at the NS7 and VP1 tree, and recombination was detected only among genotypes from the same clusters. Similarly, the GII polymerase types showed one large and two small clusters, and recombination was restricted to capsid genotypes linked to the viruses grouped in the same NS7 cluster. One of the smaller groups included viruses presenting the GII.P6, GII.P7, GII.P8, GII.P20, and GII.P36 polymerase that were linked only to the GII.6, GII.7, GII.8, GII.9, GII.14, or GII.20 capsid genotypes, while the other small group included recently described genotypes (GII.22-GII.27) and polymerase types (GII.P22-GII.P27, GII.P40). The latter could be embedded within the larger group that include all other GII viruses because of the phylogenetic clustering and a limited number of viruses from three genotypes (GII.2, GII.5, and GII.12) presenting the newly described polymerase types (Fig 8) . Notably, GII.P15 and GII.P28 did not follow the recombination restriction, as they were associated with only the capsid from non-GII viruses, GIX.1 and GVIII.1, respectively. Viruses presenting these combinations confirm the possibility of inter-genogroup recombination; however, no parental viruses have been detected so far. GI and GII phylogenetic trees from each of the non-structural protein-encoding sequences showed similar topologies (S8 Fig). In both genogroups, the NS7 clusters (color-shaded in S8 Fig) were preserved in other non-structural proteins, suggesting a limitation for recombination within the ORF1. The blue cluster in GII was nested in the light blue cluster in the NS6 phylogenetic tree, but still presented distinct separation from each other. Those clusters were disrupted in the VP1-and VP2-based trees in GII (S8 Fig) . Together, NS7 (or ORF1-based) clusters were associated with the restriction and pattern of recombination (recombination group). To determine the mechanisms associated with the restrictions on the observed ORF1/ORF2 recombination, we examined the diversification pattern of different regions adjacent to the ORF1/ORF2 junction region. Thus, we examined 200 nt from the 3'-end of the NS7 (ORF1) and 200 nt from the 5'-end of the VP1 (ORF2). This junction region is highly conserved [27] , particularly within the~20 nt encompassing the ORF1/2 overlapping region ( Fig 9A) . To explore the sites correlated with groups presenting recombination restriction, we performed a multidimensional scaling (MDS) analysis using nucleotide sequence variation at the ORF1/2 junction (Fig 9B) . The 400 nt junction region was divided into four windows, each corresponding to 100 nt in length. The first two regions correspond to the 3'-end of ORF1, and the last two correspond to 5'-end of ORF2 sequences (Fig 9A) . These MDS maps indicated strong clustering in the first window (Fig 9B) , while the second window presented modest clustering by recombination group. On the other hand, the two windows from the 5'-end of ORF2 showed strong clustering by genogroups, but not by recombination groups. The MDS maps were also created from genome regions by their function: namely 3'end of ORF1 (NS7), 5'end of ORF2 (sub-genomic RNA), predicted stem-loop, and linker sequence between predicted Genotypes were determined for all the viruses in the nearly full-length dataset using Norovirus Genotyping Tool [102] . Viruses were grouped phylogenetically and the number of viruses with a given capsid and polymerase types was recorded in each cell. Phylogenetic trees were calculated using a subsampled dataset: a maximum of two viruses from each combination of genotype and polymerase type. The colored boxes in the matrix indicate the recombination groups associated with the phylogenetic clustering on the NS7-encoding nucleotide sequences. The crossing lines on the phylogenetic tree of the VP1 indicate the disruption of VP1-clustering and recombination groups. https://doi.org/10.1371/journal.ppat.1009744.g008 Intra-genotype recombination requires mixed-infections in the host followed by co-infection of cells [28, 61] . One advantage of the full-length sequencing platform presented here is that we utilized full-length PCR amplicons and next-generation sequencing (NGS) technology [14] , which enabled us to detect reads from different viruses in a given sample. We found five mixed-infection cases in our virus collection, three from cross-sectional studies and two from cohort studies. The three cases from the cross-sectional study included mixed-infections with GII.14 and GII.4, GII.2 and GII.14, and GII.12 and GII.6 ( Fig 10A) . By applying a clonal population analysis, which determines the clones (haplotypes) present in the sample by association of NGS reads by similarity [62,63], we reconstructed the consensus sequences from multiple clones or genotypes at the near full-length level (>7400 nt, Fig 10A) . Two cases from Peruvian birth cohort and family cohort studies included mixed-infection with GII.4 and GII.3, and GII.NA2 and GII.4 viruses. In these studies, stool samples were collected regularly, which allowed us to follow up virus diversification during the shedding phase of these two reinfection cases (Fig 10B and 10C) . Child NV066X was infected with GII. Fig 10C) . On Day 0, most of the genomic reads were mapped against GII.4 virus (86% in total reads), while GII.NA2 was detected as a very minor population (6%). This prevalence was reverted by Day 15, with GII. NA2 becoming predominant on Day 15 and 21. Finally, GII.4 virus disappeared and only GII. NA2 virus was detected on Day 29. Genome titers calculated by qPCR recapitulated this trend (S10A Fig). The presence of mucosal IgA was measured using the stool samples and norovirus virus-like particles (VLPs) as antigens. Surprisingly, norovirus genotype-specific IgA response was detected only against one of the infecting genotypes, the major clone at the end of the infection, and not the one first cleared (S10B Fig) . Viral mutants (subclones) were detected in both major viruses at the end of shedding periods, possibly as a result from the mucosal response (Fig 10B and 10C) . Importantly, no chimeric sequences (i.e. indication of recombinant genomes) were detected in any of the five cases described above. The NGS reads were clearly mapped against two reference genomes, and none of them were mapped across the two different reference sequences (Fig 10) . No recombinant genomes were detected during the prolonged shedding either, which account for up to 4 weeks of mixed infections. Chimeric reads were further searched for those prolonged shedding cases using artificial chimeric genomes as references -sequence regions from two reference genomes were artificially exchanged at the ORF1/2 junction-to enforce the mapping tool to detect chimeric reads with no positive results (S11 and S12 Figs). Another evidence for the limited intra-genotype recombination of human noroviruses is revealed by tracking the changes in the polymerase type along the evolutionary trees of VP1 (Fig 11) . Examination of the phylogenetic trees from four of the most frequent norovirus genotypes reconfirmed that most recombination events occurred with polymerase types from the same branches on the NS7 phylogenetic tree (Fig 11) . The number of changes of polymerase types along branches, as counted with Markov-jump counting method [64], suggested limited recombination events during the evolution of these GII genotypes during the last 40 years. Thus, GII.2 viruses present two major polymerase types, GII.P2 and GII.P16, and recombination from GII.P2 to GII.P16 occurred three times throughout their evolution while recombination with other minor polymerase types occurred only once. In GII.6 viruses, recombination from GII.P6 to GII.P7, which are very close to each other, also occurred three times but not vice versa. The GII.4 viruses circulating in the 1970s and early 1980s presented the GII. (Fig 11) . No opposite direction of recombination (from GII.P16 to GII.P4 or GII.P31) was estimated from the tree. Finally, GII.3 viruses experienced multiple recombination events: three times from GII.P21 to GII.P16, two times from GII.P21 to GII.P12 and GII.P41 to GII.P3, and a single recombination event was recorded for GII.P3 to GII.P16, GII.P41 to GII.P29, GII.P29 to GI.P21, and GII.P12 to GII.P16. While GII.2 and GII.3 present over ten recombination events, that still could be regarded as a low number of recombination considering that this occurred over four decades. Together, this shows that recombination in human noroviruses is not a frequent event, or a small number of recombinant viruses are fit to emerge and circulate in the human population. Because of the public health impact and development of new technologies to study virus genomics, the epidemiology and diversification of norovirus has been widely investigated [13, 14, 66] . Over the last two decades, a great number of norovirus sequences have been recorded in public databases; however, this public dataset is currently biased at the geographical and temporal level, and only limited genotypes and genomic regions have been widely sequenced. Thus, most records belong to capsid sequences from the major genotypes (GII.2, GII.3, GII.4, GII.6, and GII.17) detected in developed countries since the 2000s. In this study, we successfully closed some of the information gaps by adding new sequence information for 24 genotypes and over 80 noroviruses circulating prior to 2000. Removal of temporal bias improved the fitness of the molecular clock analyses and our estimates of norovirus evolution. Many studies have reported the evolutionary rate of noroviruses, with some suggesting that differences in the evolutionary rate among different norovirus genotypes could account for the emergence and predominance of certain viruses [21, 67] . Our analyses showed that norovirus presented similar rates of evolution across the different ORFs and genotypes/variants (mean 1.37-5.38 × 10 −3 substitutions/site/year) with overlapping 95%HPD intervals. Some genotypes, e.g., GII.1, GII.3, and GII.6, presented substitution rates that did not show overlapping of the 95%HPD intervals as compared with other genotypes; however, those differences were not related with the epidemic potential of the viruses. Thus, newly emerged or predominant viruses, like GII.2, GII.6, or GII.17 did not present higher rates of evolution as compared to other minor genotypes. Diversifying pressure on VP1, mostly on the P2 sub-domain, was detected in predominant genotypes and those with multiple variants. However, only viruses from the GII.4 genotype accumulate changes on the VP1 protein. Detection of diversifying pressure on non-GII.4 viruses with multiple variants suggests that these different lineages emerged, rapidly adapted to the human population, but showed low genetic robustness to accommodate major changes on their VP1 protein after decades of circulation and evolution. The significance of the co-circulation of multiple non-GII.4 variants is not completely understood. Changes in susceptibility and antigenicity have been attributed to the recent emergence of the new GII.17 variant that predominated during 2014-2015 in Asia [25, 26] ; however, additional studies are required to determine the phenotypic differences that could account for the emergence and co-circulation of the other non-GII.4 variants. Similar to other viruses, emergence of new norovirus is probably marked by three evolutionary steps: (i) acquisition of mutations that would provide an initial advantage, (ii) subsequent mutations that would result in the proper adaptation to epidemic potential, and (iii) dispersion of the virus resulting in various mutations that follow a stochastic pattern [16, 68] . These series of events have been shown to occur in the emergence of different GII.4 variants [14, 16, 69] , but most recently in the emergence of predominant GII.17 viruses. With stronger surveillance systems, the emergence of GII.17 was characterized by the quick change from one transiently circulating variant (namely C, circulated during 2012-2015) to the predominant variant (namely D, which was detected since 2013 and is still reported in certain countries, e.g. Japan [70] ). Thus, this rapid adaptation to the human population makes the detection of the intermediate viruses very difficult. In our historical samples, by focusing the attention on those less prevalent viruses, we detected multiple viruses that branched between defined variants or genotypes. While different hypotheses were presented on the emergence of novel norovirus (e.g. spill-over from hidden animal reservoirs [71] or variants originated in immunocompromised individuals [72] ), it is possible that the genetic drift operates over long periods and the intermediate viruses are cryptically circulating in under-sampled populations, as shown by the recent detection of 'novel' noroviruses in active surveillance in communities [44,45,73] or country-wide monitoring systems [51] . Considering the cryptic genetic drift and rapid adaptation to achieve epidemic potential, we may need to accommodate multiple evolutionary steps, i.e. genetic drift, rapid adaptation, and exponential growth of viral population followed by predominance in the communities, to the evolutionary models or in-depth analyses of the intrahost evolution of noroviruses to fully understand the emergence of novel viruses. A paradigm of norovirus epidemiology is that changes on the VP1 protein will prompt the emergence of new noroviruses by facilitating the escape from herd immunity [47] . While this seems to be the case for GII.4 noroviruses that presented continuous changes on the VP1, the recent predominance of GII.2 has been explained by changes on the viral polymerase [23] . Thus, it has been suggested that higher mutational rates, escape from T-cell immune responses to non-structural proteins, or increases in the replicative efficiency that results in higher levels of virus shedding could enhance transmissibility [33, 35, 69, 74] . In that regard, the role of other each polymerase type was determined from the phylogeny of the NS7-encoding nucleotide sequences. The direction and number of changes of polymerase types, i.e. recombination events, were estimated using Markov jump counting along the branches. Events with mean frequency >0.5 were summarized as bar graphs (the mean and the 95% highest posterior density interval). https://doi.org/10.1371/journal.ppat.1009744.g011 non-structural proteins has been largely overlooked. Thus, in addition to mutations on the viral polymerase (NS7), our study found high variability and mutations on the NS1/2 (Nterm) and NS4 (3A-like) proteins from epidemic vs. endemic GII.P16 and GII.P31 viruses. NS1/2 and NS4 proteins have been shown to induce replication complex formation by recruiting cellular membranes, and NS1/2 has been suggested to mediate and control the pathways of innate immunity [75, 76] . Whether changes on these proteins are the consequence of founder effect or indeed enhance the epidemic potential from some viruses, warrants further research. Another paradigm of norovirus epidemiology and evolution is the frequent interchange of genomic regions, by means of recombination, that results in new viruses with epidemic potential [12, 27, 28] . A hot spot for norovirus recombination has been identified at the ORF1/ORF2 overlapping region [27] , reinforcing the role of non-structural proteins in the emergence of new noroviruses. Although a very large number of combinations of genotypes could be present in nature by means of recombination, our findings indicate that noroviruses show restriction in their ability to recombine with phylogenetically unrelated genotypes. This was supported by large-scale population genomics analyses, as well as the analysis of the intra-host diversity of patients with mixed infections. The latter provided a unique opportunity to determine the frequency of recombination among two viruses from the same phylogenetic cluster (i.e. GII. While one chimeric read (0.005% of coverage) was detected from the second sample (day 7), this was not reproducible in a second experiment. These data indicate that recombination at the ORF1/2 overlapping regions may occur under strong restriction and less frequently than initially thought. Previous studies reported a small number of recombinant viruses with capsid and polymerase combinations that do not follow the proposed restriction in this study, e.g. GII. 4[P7] or GII.7[P21] [28, 30] . Careful interpretation should be given to those reports, as the recombinants were determined with short-length sequences that could be incorrectly typed and/or amplicons from separated regions of the genomes that could be an experimental artifact of mixed infections. Whether recombination events occur more frequently at the intragenotype or at the variant level [41,50] remains to be determined. A limitation of our analyses is that it is technically difficult to define recombinants or false-positives among genetically similar viruses with intermediate viruses undetected. Template switching between genomic and sub-genomic RNA is widely accepted as the molecular mechanism for ORF1/2 recombination in noroviruses [3, 28] ; however, there are no sufficient data to determine whether the sub-genomic or genomic RNA works as the acceptor molecule. Simmonds et al. predicted a stem-loop structure at the 3' end of ORF1 (�7 nt downstream of the (-) sub-genomic RNA start position) [60] , that was shown to be part of the subgenomic RNA promoter in murine norovirus [77] . If template switching occurs at the stemloop as suggested in the copy-choice model [78] , acceptor genome should be derived from genomic RNA with this stem-loop promoter sequence. Indeed, Bull et al. estimated the recombination breakpoints were located on average 16-19 nt upstream from the start of (+) ORF1/ ORF2 overlap [29] . Our finding supports the idea that genomic RNA, not sub-genomic RNA, works as an acceptor template. The acceptor genomic RNA could be annealed with the donor template at the 3' end of the ORF1, and thus the template-switching is restricted by the sequence identity on those regions between donor and acceptor RNA molecules. Based on these recombination restrictions, human noroviruses seem to be clustered into five recombination groups of viruses that do not genetically interact among each other. Restrictions on recombination of genes have been shown for multiple positive-sense RNA viruses, including enterovirus [79] and flavivirus [80] . Additionally, restrictions on the reassortment of genes have been reported for segmented viruses such as influenza virus and rotavirus [81, 82] . These restrictions seem to be governed by protein-protein and protein-nucleic acid interactions [83] [84] [85] [86] [87] [88] . This study supports the mechanism of restriction to nucleic acid-nucleic acid interactions during the processing of (-) RNA genomes. Additional studies are warranted to determine whether exchange of genomes among the different genotypes may provide fitness disadvantages [89] [90] [91] , by which certain proteins interact better with the structural proteins or with host machinery during the replication process. Also, it is to be studied whether these restrictions occur in other positive-sense RNA viruses that utilize recombination as a mechanism of diversification [79, 92] . Development of a simple norovirus genome sequencing protocol provided us the opportunity to analyze archival samples and different aspects of norovirus diversification and evolution. In conclusion, we showed that human noroviruses present (i) a similar rate of evolution at all genomic regions and genotype/variant levels, (ii) intermediate evolutionary states that might be a key for the adaptation to emerging viruses, (iii) mutations on non-structural proteins that could provide novel characteristics, and (iv) restrictions on the ability to recombine different (ORF1/ORF2) genomic regions, which results in co-circulating populations of viruses evolving independently. Moreover, the new sequence information and analyses reported here could provide baseline information for the study of future epidemics and ultimately the prevention of norovirus infection. Studies were originally approved or exempted by IRB from each respective institution. Archival stool samples stored at different laboratories were analyzed anonymously and collectively exempted under FDA institutional IRB (16-069B). We retrospectively analyzed archived fecal samples positive for human norovirus, focusing on sequencing minor norovirus genotypes or historical viruses collected in the 1970s, 1980s, 1990s, and early 2000s (S1 Table) . The samples were collected as part of studies conducted in nine laboratories [52,56, [93] [94] [95] [96] [97] [98] [99] and the World Health Organization in different countries during 1976-1979 [100, 101] . Nearly full-length deep sequencing of archival viruses was retrospectively performed as previously described [14] . Briefly, complementary DNA (cDNA) was generated from the extracted viral RNA genome using the Maxima Minus First Strand cDNA Synthesis Kit (ThermoFisher Scientific) and a poly-A primer. The full-length viral genome amplification was done using the SequalPrep Long PCR Kit (Invitrogen) and primers listed in S3 Table. The resulting full-length viral genome amplicons (~7.5 kb) were run on the 1% agarose gel and extracted using Qiagen Gel Extraction Kit (Qiagen, California, USA). The gel-extracted amplicons were quantified using the Qubit dsDNA HS Assay Kit (ThermoFisher Scientific), and subjected to NGS using MiSeq system (illumina, California, USA). The library for NGS was prepared using the Nextera XT DNA Library Prep Kit (illumina), and the paired-end 2 × 250 bp sequence reads were obtained. Reads were quality-filtered (base quality score � 20, and depth of coverage � 10) and i) mapped against reference genome set (all genotypes and polymerase types implemented in Norovirus Genotyping Tool [102] ) to screen the amplified genotypes, and ii) mapped against corresponding full-or nearly full-length reference genomes to reconstruct its consensus sequence using HIVE platform [62] . Samples were re-processed from the 10% stool suspension to confirm the mixed-infection if multiple genotypes were detected. Reads from mixed infection samples were then mapped against corresponding reference genomes followed by clonal population analysis using the Hexahedron Coordinated Clonal Analysis Tool implemented in the HIVE platform [63] . This clonal population analysis separately assembled all possible viral population (clones) contained in a single sample, and reconstructed their own consensus genomes at a near full-length level. Finally, chimeric reads were searched by HIVE using artificial chimeric genomes (400 nt ORF1/2 junction region) as references, and readmapping against the references were visualized with MSAViewer [103] . Nearly-complete consensus genome sequences and raw NGS reads from shedding cases obtained in this study were deposited in GenBank (Accession numbers: MG706448, MK733201-MK733207, MW261787-MW261800, MW305481-MW305742; summarized in S1 Table) and SRA (BioProject accession number: PRJNA659534), respectively. Along with our own genome set, we collected the nearly-full length (�7000 nt) and/or VP1-encoding (�1500 nt) GI, GII, GVIII, and GIX sequences from GenBank (as of March 30, 2020, S4 and S5 Tables). Sequences from animals, environment (e.g. sewage water), and immunocompromised patients were excluded. All the genomes were genotyped using Norovirus Genotyping Tool [102] . Collection years of the sequences were summarized using R v3.6.0 and GraphPad Prism v7. Geographic location (countries) were summarized and visualized using maptools and mapplots packages in R. Map shape file (1:10m Cultural Vector, Admin 0 -Countries) was obtained from Natural Earth website (https://www.naturalearthdata.com/ downloads/10m-cultural-vectors/10m-admin-0-countries/, accessed on June 25, 2021). To avoid false phylogenetic signals arising from genome variability and/or recombination events, sequences were split into genogroups/genotypes and each genomic region in the encoding protein (NS1/2-7, VP1, and VP2). Sequences were then multiple-aligned using MUSCLE [104] , MAFFT (for large dataset) [105] , or TranslatorX (for protein-coding codon sequence alignment) [106] . Maximum-likelihood (ML) phylogenetic trees were built using PhyML for nucleotide and amino acid sequences, with best-fit substitution models estimated with Smart Model Selection and branch support estimated with approximate likelihood-ratio test implemented in PhyML [107] . Variants of each genotype were determined based on the cutoff of the 5% difference on the VP1 amino acid sequence and/or the proposed classification criteria using standard deviation of inter-/intra-variant patristic distance on VP1-based ML trees [11] using R and ape package. Evolutionary patterns, i.e. clock-likeness, rate of evolution, and selective pressure were analyzed with ORF2 (VP1-encoding) sequences. Clock-likeness of the dataset (genotypes/variants) was confirmed with linear regression analysis of root-to-tip distance using ML trees and TempEst v1.5 [108] , with and without historical sequences. The rate of evolution (substitutions/site/year) was calculated using BEAST v1.10.4 with strict-clock model [109] . The SRD06 model was used for estimating the nucleotide substitution process of the VP1-encoding sequences. The population size was assumed to be constant throughout their evolutionary history. The first 10% of the logs from the Markov chain Monte Carlo runs were removed as a burn-in before summarizing the posterior values. Recombination frequency, i.e. change of polymerase type, throughout the evolutionary history was estimated using asymmetric Markov Jump counting method [64] implemented in BEAST v1.10.4. The amino acid accumulation over time (distance from the oldest virus within each genotype) was calculated using phangorn package and R, with JTT model as the best-fit amino acid substitution model (in most of the genotypes) estimated using Smart Model Selection. Site-by-site selective pressure on VP1 from each genotype was estimated using MEME (Mixed Effects Model of Evolution) that detects sites under episodic selection by assuming varying pressure across branches on the phylogenetic tree [110] and FUBAR (Fast, Unconstrained Bayesian AppRoximation) that assumes constant pressure on the entire tree [111] , using Datamonkey (small dataset) [112] or HyPhy v2.5.14 (large dataset with �500 sequences) [113] . Statistically significant sites (P<0.05 with empirical Bayes Factor on internal branches>100 in MEME and Bayes Factor>0.9 in FUBAR) were plotted using GraphPad Prism 7. In order to reduce the sampling bias derived from recent large-amount sequence submissions of GII.2, GII.4, and GII.17 viruses, we generated subsampled dataset which includes a maximum of 20-30 randomly selected sequences per variant per year (n = 302 for GII.2, n = 821 for GII.4, and n = 204 for GII.17 viruses), and used for root-to-tip regression, evolutionary rate, and selective pressure analyses. To overview the evolutionary pattern of other proteins, the genome-wide nucleotide and amino acid sequence diversity was calculated as site-by-site Shannon entropy values using Entropy-One Tool (https://www.hiv.lanl.gov/content/sequence/ENTROPY/entropy_one. html). The amino acid accumulation pattern was also estimated for other nonstructural proteins and VP2 proteins from each genotype. Maximum 50 nt of the 5' or 3' end of the NS1/2 or VP2-encoding region, respectively, was removed from the analyses to exclude the sequence gaps derived from partial genomes. To account for the sampling bias by predominance of certain genotypes, the ML phylogenetic trees and Shannon entropy values for each genomic region was estimated using randomly subsampled dataset which included a maximum of two viruses from each polymerase and capsid genotype combination (n = 33 for GI and n = 119 for GII viruses). The evolutionary rate of ORF1 (NS1/2-NS7) and ORF3 (VP2) sequences were estimated using BEAST v1.10.4 as described above. The ORF1 sequences were partitioned by each non-structural protein and rate of evolution was jointly estimated. The amino acid distance from the oldest virus within each genotype was calculated for the NS4, NS7, and VP2 proteins using R as described above. To explore key amino acid residues on nonstructural and VP2 proteins that could be associated with global spread and the epidemic potential of noroviruses, phylogenetic analyses were conducted using the amino acid sequences from NS1/2-NS7 and VP2 from GII.P4, GII.P16, GII.P31, and GII.2 viruses. The ML phylogenetic trees were built using PhyML as indicated above, and ancestral sequences on the nodes that diverged into epidemic clusters were estimated using maximum likelihood method implemented in phangorn package in R. The amino acid mutations on the NS7 of epidemic viruses were mapped on the structural model of the polymerase from GII.P4 virus (PDB number 4QPX) using UCSF Chimera v 1.11. Recombination combination was tabulated using the nearly full-length sequence dataset with their polymerase and capsid genotype information, and visualized as a heatmap using GraphPad Prism v7. The 400 nt sequence spanning the ORF1/2 junction region was extracted from the nearly fulllength sequence dataset. The Shannon entropy values on this junction region were calculated using Entropy-One Tool as described above. In order to determine the restriction factors on recombination, we explored the 2D-MDS map generated from pairwise nucleotide differences on the junction region. The junction region was split into four non-overlapping windows (1-100, 101-200, 201-300 , and 301-400 nt) and/or functional regions (3' end ORF1, 5' end subgenomic region, predicted stem-loop, and linker sequence between predicted sub-genomic promoter and sub-genomic RNA [60]), and corresponding 2D-MDS maps were generated using R and stat package. We detected two mixed infection cases from birth cohort [93] and family cohort studies conducted in Lima, Peru. By following up those two cases, we successfully observed 20-days or 29-days prolonged shedding of mixed infection with two different genotypes: one child with GII.3 and GII.4, and another with GII.4 and GII.NA2, respectively. In addition to NGS and subsequent clonal population analysis, we performed genotype-specific qPCR to quantify the viral load of each genotype during the shedding. Briefly, viral RNA was quantified by duplex one-step real-time PCR with GII-specific primers and genotype-specific probes (S6 Table) using TaqMan Fast Virus 1-Step Master Mix kit (Applied Biosystems). The genotype-specific control plasmids were generated in house with corresponding partial viral genomes inserted into pCI vectors (Promega). In addition, with the 10% PBS suspension of the stool samples, we detected human IgA response with corresponding VLPs as antigens by ELISA. The VLPs were produced as described elsewhere using baculovirus expression system [8, 16, 114] with VP1-encoding sequences from GII.3 (Maizuru010524; accession number EF547399), GII.4 (Rock-villeD1; KY424328), and GII.NA2 (PNV06929; MG706448) viruses. Aetiology-Specific Estimates of the Global and Regional Incidence and Mortality of Diarrhoeal Diseases Commonly Transmitted through Food Systematic literature review of role of noroviruses in sporadic gastroenteritis Human norovirus transmission and evolution in a changing world Visualization by immune electron microscopy of a 27-nm particle associated with acute infectious nonbacterial gastroenteritis X-ray crystallographic structure of the Norwalk virus capsid Atomic resolution structural characterization of recognition of histo-blood group antigens by Norwalk virus Human Monoclonal Antibodies That Neutralize Pandemic GII.4 Noroviruses Genotype-Specific Neutralization of Norovirus is Mediated by Antibodies Against the Protruding Domain of the Major Capsid Protein The 3' end of Norwalk virus mRNA contains determinants that regulate the expression and stability of the viral capsid protein VP1: a novel function for the VP2 protein Understanding the relationship between norovirus diversity and immunity Emergence of norovirus strains: A tale of two genes Molecular surveillance of norovirus, 2005-16: an epidemiological analysis of data collected from the NoroNet network Static and Evolving Norovirus Genotypes: Implications for Epidemiology and Immunity Antigenic Cartography reveals complexities of genetic determinants that lead to antigenic differences among pandemic GII.4 noroviruses Population Genomics of GII.4 Noroviruses Reveal Complex Diversification and New Antigenic Sites Involved in the Emergence of Pandemic Strains Steep rise in norovirus cases and emergence of a new recombinant strain GII Norovirus GII.P16/GII.2-Associated Gastroenteritis, China Genome of Emerging Norovirus GII in Japan reveal a novel polymerase sequence and amino acid substitutions in the capsid region Rapid emergence and predominance of a broadly recognizing and fast-evolving norovirus GII.17 variant in late 2014 Characterization of blockade antibody responses in GII.2.1976 Snow Mountain virus-infected subjects Phylogenetic Analyses Suggest that Factors Other Than the Capsid Protein Play a Role in the Epidemic Potential of GII.2 Norovirus. mSphere Characterization of the new GII.17 norovirus variant that emerged recently as the predominant strain in China An outbreak caused by GII.17 norovirus with a wide spectrum of HBGA-associated susceptibility Emergence of Novel Human Norovirus GII.17 Strains Correlates With Changes in Blockade Antibody Epitopes Norovirus recombination in ORF1/ORF2 overlap Norovirus recombinants: recurrent in the field, recalcitrant in the lab-a scoping review of recombination and recombinant types of noroviruses Norovirus recombination Long intervals of stasis punctuated by bursts of positive selection in the seasonal evolution of influenza A virus Evolution of norovirus Japan: National Institute of Infectious Diseases Animals as Reservoir for Human Norovirus What is the reservoir of emergent human norovirus strains? Norovirus Infection and Disease in an Ecuadorian Birth Cohort: Association of Certain Norovirus Genotypes With Host FUT2 Secretor Status Higher Viral Load of Emerging Norovirus GII.P16-GII.2 than Pandemic GII.4 and Epidemic GII A Secreted Viral Nonstructural Protein Determines Intestinal Norovirus Pathogenesis Membrane alterations induced by nonstructural proteins of human norovirus The murine norovirus core subgenomic RNA promoter consists of a stable stem-loop that can direct accurate initiation of RNA synthesis Why do RNA viruses recombine? Largescale genomic analysis reveals recurrent patterns of intertypic recombination in human enteroviruses The extent of homologous recombination in members of the genus Flavivirus Evolutionary dynamics of human rotaviruses: balancing reassortment with preferred genome constellations The evolutionary genetics and emergence of avian influenza viruses in wild birds Critical role of segment-specific packaging signals in genetic reassortment of influenza A viruses Genetic determinants restricting the reassortment of heterologous NSP2 genes into the simian rotavirus SA11 genome Predicting Intraserotypic Recombination in Enterovirus 71 Group A human rotavirus genomics: evidence that gene constellations are influenced by viral protein interactions H5N8 and H7N9 packaging signals constrain HA reassortment with a seasonal H3N2 influenza A virus Compatibility among polymerase subunit proteins is a restricting factor in reassortment between equine H7N7 and human H3N2 influenza viruses Experimental evidence of recombination in murine noroviruses Replicative fitness recuperation of a recombinant murine norovirus-in vitro reciprocity of genetic shift and drift Rules governing genetic exchanges among viral types from different Enterovirus A clusters Origin and evolution of pathogenic coronaviruses Multiple norovirus infections in a birth cohort in a Peruvian Periurban community Wide variety of recombinant strains of norovirus GII in pediatric patients hospitalized with acute gastroenteritis in Thailand during A predominant role for Norwalk-like viruses as agents of epidemic gastroenteritis in Maryland nursing homes for the elderly Molecular epidemiology of norovirus outbreaks in Argentina Molecular characterization of calicivirus strains detected in outbreaks of gastroenteritis in Argentina Molecular epidemiology of norovirus strains in Paraguayan children during 2004-2005: description of a possible new GII.4 cluster Molecular characterization of calicivirus strains detected in outbreaks of gastroenteritis occurring in Argentina during Annual Report of Program Activities Epidemiology and evolution of rotaviruses and noroviruses from an archival WHO Global Study in Children (1976-79) with implications for vaccine design An automated genotyping tool for enteroviruses and noroviruses MSAViewer: interactive JavaScript visualization of multiple sequence alignments MUSCLE: multiple sequence alignment with high accuracy and high throughput MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization TranslatorX: multiple alignment of nucleotide sequences guided by amino acid translations New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0 Exploring the temporal structure of heterochronous sequences using TempEst (formerly Path-O-Gen) Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10 Detecting individual sites subject to episodic diversifying selection FUBAR: a fast, unconstrained bayesian approximation for inferring selection Datamonkey 2.0: A Modern Web Application for Characterizing Selective and Other Evolutionary Processes HyPhy 2.5-A Customizable Platform for Evolutionary Hypothesis Testing Using Phylogenies Comparative evolution of GII.3 and GII.4 norovirus over a 31-year period We thank the Facility for Biotechnology Resources at U.S. FDA for their support with next generation sequencing. We thank Karina Rivero for her technical support in Argentina. We thank the community of San Juan de Miraflores for their collaboration and Asociación Benéfica PRISMA for the fieldwork in Peru.