You are on page 1of 14

Journal of General Virology (2014), 95, 278–291 DOI 10.1099/vir.0.

059634-0

Review Norovirus gene expression and replication


Lucy G. Thorne and Ian G. Goodfellow
Correspondence Division of Virology, Department of Pathology, University of Cambridge, Addenbrooke’s Hospital,
Ian G. Goodfellow Hills Road, Cambridge CB2 2QQ, UK
Ig299@cam.ac.uk

Noroviruses are small, positive-sense RNA viruses within the family Caliciviridae, and are now
accepted widely as a major cause of acute gastroenteritis in both developed and developing
countries. Despite their impact, our understanding of the life cycle of noroviruses has lagged
behind that of other RNA viruses due to the inability to culture human noroviruses (HuNVs). Our
knowledge of norovirus biology has improved significantly over the past decade as a result of
numerous technological advances. The use of a HuNV replicon, improved biochemical and cell-
based assays, combined with the discovery of a murine norovirus capable of replication in cell
culture, has improved greatly our understanding of the molecular mechanisms of norovirus
genome translation and replication, as well as the interaction with host cell processes. In this
Received 18 September 2013 review, the current state of knowledge of the intracellular life of noroviruses is discussed with
Accepted 13 November 2013 particular emphasis on the mechanisms of viral gene expression and viral genome replication.

Introduction currently into five genera: Vesivirus, Lagovirus, Nebovirus,


Sapovirus and Norovirus. Members of the genera Norovirus
The prototype norovirus, Norwalk virus, was first described
and Sapovirus are able to infect humans and cause
in 1972 as the aetiological agent responsible for an outbreak
gastroenteritis. The genus Norovirus is subdivided into at
of acute gastroenteritis in an elementary school in Norwalk,
least five genogroups (GI–V). Genogroups GI, GII and GIV
OH, USA (Kapikian, 2000). Noroviruses are now accepted
infect humans and cause acute gastroenteritis, but
as a leading cause of gastroenteritis in developed and
noroviruses have also been isolated from numerous other
developing countries (Glass et al., 2009; Hall et al., 2013).
species including pigs (GII), cattle and sheep (GIII), and
Spread primarily via the faecal–oral route, norovirus
mice (GV). More recently, a novel norovirus identified in
infections are typically an acute self-limiting gastrointestinal
domestic dogs with diarrhoea has been proposed to
infection. Norovirus gastroenteritis has recently been
represent a new genogroup, GVI (Mesquita et al., 2010).
identified as a significant cause of morbidity and mortality
To date, zoonotic norovirus infections have not been
in the immunocompromised, and can result in long-term
reported, but there is clear evidence of the potential for
persistent disease (reviewed by Bok & Green, 2012).
transmission. For example, HuNVs can infect gnotobiotic
Norovirus infection has also been associated with a number
piglets (Cheetham et al., 2006) and there is serological
of more significant clinical outcomes: necrotizing entero-
evidence of HuNV in pigs (Farkas et al., 2005). Whilst
colitis (Turcios-Ruiz et al., 2008), seizures in infants (Medici
antibodies to GVI noroviruses have been identified in
et al., 2010), encephalopathy (Ito et al., 2006), pneumatosis
veterinarians, it has yet to be determined whether infection
intestinalis (Chan et al., 2010; Kim et al., 2011) and
results in clinical disease (Mesquita et al., 2013). Additional
disseminated intravascular coagulation (CDC, 2002), to
studies are required to further elucidate the zoonotic
name but a few. In developing countries, an estimated
potential of noroviruses and whether animals represent a
200 000 deaths in children ,5 years of age are thought to be
reservoir from which new strains may emerge.
due to norovirus infections (Patel et al., 2008) and they have
recently been reported as the second leading cause of
gastroenteritis-related deaths in the USA, typically resulting Norovirus genome organization
in 797 deaths per annum (Hall et al., 2013). Despite their
The norovirus genome is a compact, positive-sense ssRNA
significant impact, noroviruses remain one of the most
molecule, ranging in size from 7.3 to 7.5 kb across
poorly characterized groups of RNA viruses, due largely to
the genus (Fig. 1). The 59 end of the genomic RNA is
the fact that, despite numerous attempts (Duizer et al., 2004;
covalently attached to a virus-encoded protein known as
Papafragkou et al., 2013; Takanashi et al., 2013), human
VPg, whilst the 39 end is polyadenylated. The UTRs at
noroviruses (HuNVs) have yet to be cultured efficiently in
either end of norovirus genomes are typically short, e.g. the
cell culture.
59 and 39 UTRs of murine norovirus (MNV) are 5 and
Noroviruses are members of the family Caliciviridae 78 nt, respectively, and the 39 UTR of HuNV is typically
of small, positive-sense RNA viruses, which is divided 48 nt (Gutiérrez-Escolano et al., 2000a; Karst et al., 2003;
Downloaded from www.microbiologyresearch.org by
278 059634 G 2014 SGM Printed in Great Britain
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

(a) Human norovirus ORF1 ORF2 ORF3


VP1
VPg p48 NTPase p22 VPg Pro Pol VP2 AAAA(n)
N-term 2C-L 3A-L

VP1
VP2 AAAA(n)

(b) Murine norovirus


ORF1 ORF2 ORF3

VP1
NS1/2 NS3 NS4 NS5 NS6 NS7 VP2 AAAA(n)

Caspase 3
cleavage sites
VP1
VF1 VP2 AAAA(n)

Fig. 1. Organization of the HuNV and MNV genomes. (a) The HuNV genome is covalently attached to VPg (NS5) with a poly(A)
tail and is divided into three ORFs, common to all noroviruses. ORF1 is translated as a polyprotein, which is cleaved by the viral
protease NS6 to produce the NS proteins. ORF2 and ORF3 are translated from a subgenomic RNA. 2C-L, 2C-like; 3A-L, 3A-
like. (b) MNV shares a similar genome organization, but has an additional alternative fourth ORF. ORF4 overlaps with ORF2 and
is also translated primarily from the subgenomic RNA into the virulence factor 1 (VF1) protein.

Pletneva et al., 2001). The UTRs contain evolutionarily nomenclature for the norovirus proteins has yet to be
conserved RNA secondary structures that extend into the unified in the literature and is summarized in Table 1.
coding regions and can be found throughout the genome
(Simmonds et al., 2008). These structures are important for
viral replication, translation and pathogenesis (Bailey et al., Model systems for the study of norovirus gene
2010b; McFadden et al., 2011; Simmonds et al., 2008). expression and replication

The norovirus genome is organized into three conserved As highlighted above, the inability of HuNV to be cultured
ORFs (Fig. 1a), with the exception of MNV, which has a in vitro has hampered the characterization of the viral life
cycle. Prior to the discovery of MNV in 2003, numerous
fourth alternative ORF (Fig. 1b). The fourth ORF is unique
other members of the family Caliciviridae were used widely
to the MNV cluster in GV and has not yet been identified in
as model systems with which to study norovirus biology
any other norovirus; the only member of the family
(reviewed by Vashist et al., 2009). Feline calicivirus (FCV),
Caliciviridae known to have an equivalent fourth ORF is
a member of the genus Vesivirus, was the first calicivirus for
human sapovirus (Clarke & Lambden, 2000; McFadden
which a cell culture and reverse-genetics system was
et al., 2011). For all noroviruses, ORF1 is translated as a large
available (Sosnovtsev & Green, 1995), and has been used
polyprotein, which is co- and post-translationally cleaved
widely to study various aspects of the calicivirus life cycle.
by the virus-encoded protease (NS6) to release at least
More recently, however, studies have indicated that HuNV
six mature non-structural (NS) proteins, including NS6
RNA isolated from feacal samples is infectious in cell
(Sosnovtsev et al., 2006) (Fig. 1). The other NS proteins
culture, as transfection into cells results in RNA replication,
include the viral RNA-dependent RNA polymerase (RdRp;
viral protein production and release of virions into the cell
NS7), VPg (NS5), the putative NTPase/RNA helicase (NS3),
supernatant (Guix et al., 2007). Importantly, this experi-
and NS1/2 and NS4, which have both been implicated in
mental system does not allow multicycle replication as the
replication complex formation (Hyde & Mackenzie, 2010;
virus generated is unable to reinfect the neighbouring cells.
Hyde et al., 2009). NS1/2 is further processed during the
latter stages of virus infection by cellular caspases activated A number of significant steps in the understanding of
by apoptosis and an as-yet-unidentified cellular protease norovirus gene expression and replication have also been
(Sosnovtsev et al., 2006). ORF2 and ORF3 are translated made using MNV, a GV norovirus initially identified as a
from a subgenomic RNA (Fig. 1), and encode the major and lethal infection in immunocompromised mice (Karst et al.,
minor capsid proteins, VP1 and VP2, respectively. The 2003). MNV replicates efficiently in primary or immorta-
subgenomic RNA is identical to the last 2.4 kb of the lized murine dendritic and macrophage cells (Wobus et al.,
genome and is also attached covalently to VPg at the 59 end 2004), and has a number of reverse-genetics systems
with a poly(A) tail at the 39 end (Herbert et al., 1997). ORF4 available (Chaudhry et al., 2007; Ward et al., 2007; Yunus
of MNV is translated from an alternative reading frame et al., 2010). The development of a Norwalk virus replicon
within ORF2 and encodes the recently identified protein system (HuNV), where cells stably maintain Norwalk
virulence factor 1 (VF1), a mitochondrially localized novel virus RNA containing an antibiotic selection marker in
innate immune regulator (McFadden et al., 2011). The the capsid-coding region (Chang et al., 2006), added
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 279
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

Table 1. Nomenclature for HuNV and MNV proteins and their functions

MNV HuNV* Function

NS1/2 p48 (N-term) Replication complex formationD, contributes to persistence in MNV infections
NS3 NTPase (2C-like) RNA helicaseD/NTPase
NS4 p22 (3A-like) Replication complex formationD
NS5 VPg Genome-linked protein involved in translation and replication
NS6 Pro (3C-like) Protease
NS7 Pol/3Dpol RdRp
VP1 VP1 Major capsid protein
VP2 VP2 Minor capsid protein
VF1 No equivalent Virulence factor

*Alternative names for the HuNV proteins are given in parentheses, and originate through comparison of the norovirus proteins and genome
organization with those of poliovirus.
DThese functions have been proposed, but have yet to be demonstrated fully.

an additional experimental system. Biochemical studies (e.g. template for the ‘pioneer round’ of viral RNA translation
Belliot et al., 2005) and other cell-based approaches reliant (Fig. 2). For this to occur, the viral RNA is recognized by
on the expression of viral proteins in immortalized cells cellular translation initiation factors and is translated into
(Subba-Reddy et al., 2011, 2012) have also been used to gain protein by the cellular translational apparatus. RNA viruses
insights into mechanisms of viral genome replication. A typically utilize novel mechanisms for viral genome
discussion of the merits of each experimental system has translation, not only for the initiation of translation, but
been provided previously (Vashist et al., 2009). The data also to increase the coding capacity of their relatively short
reviewed below rely on a combination of studies using all genomes (Firth & Brierley, 2012) and this is also true also
these approaches and, where appropriate, the experimental for noroviruses. Translation of all calicivirus genomes
system with which the data were obtained is highlighted. occurs by a novel mechanism involving VPg that is not
found in many other animal RNA viruses (Herbert et al.,
1997). However, recent data would indicate that a similar
Norovirus binding and entry mechanisms mechanism may be employed by members of the family
The interaction of noroviruses with the cell surface is Astroviridae, for which the linkage of VPg to viral RNA is
known to involve carbohydrate structures, which in the essential for RNA infectivity (Fuentes et al., 2012), as also
case of HuNVs include the histo blood group antigens described for noroviruses (Chaudhry et al., 2006). Some
(HBGAs). A detailed description of the interaction of aspects of the mechanism of calicivirus RNA genome
HuNVs with the cell surface and HBGAs, as well as the role translation are similar to that used by the plant viruses in
this interaction plays in norovirus entry and susceptibility, the family Potyviridae (Léonard et al., 2000). VPg attached
has been provided in detail recently (Donaldson et al., covalently to the 59 end of the genome mediates translation
2008; reviewed in Donaldson et al., 2010). MNV, currently of viral RNA, acting as a cap substitute, and recruits host
cell translation initiation factors (Fig. 3). The structure of
the only norovirus that replicates efficiently in cell culture,
the MNV VPg protein has recently been described as
binds host cells via sialic acid moieties, glycolipids and
containing a compact helical core, flanked by more flexible
glycoproteins in a strain-dependent manner (Taube et al.,
intrinsically disordered N- and C-terminal regions (Leen
2009, 2012) The mechanism by which MNV enters cells has
et al., 2013). This high degree of flexibility is likely to be
yet to be elucidated fully, but it is dependent on dynamin
due to the numerous roles that VPg plays in the norovirus
and cholesterol (Gerondopoulos et al., 2010; Perry &
life cycle (Goodfellow, 2011). The VPg proteins of HuNV
Wobus, 2010). As mentioned above, receptor binding and/
and MNV have both been shown to interact with
or the entry process appear to be at least one of the limiting
components of the eIF4F translation initiation factor
factors for the culture of HuNV in immortalized cells as complex, in particular eIF4E (Chaudhry et al., 2006;
viral RNA transfected into cells is able to undergo limited Goodfellow et al., 2005), the cap-binding protein, and
replication. Additional work in this area may therefore eIF3 (Daughenbaugh et al., 2003), which is recruited to the
provide a mechanism with which to further understand the complex and in turn helps to recruit the 43S ribosomal
block to HuNV replication in cell culture. pre-initiation complex (Fig. 3). Functional studies have to
date only been possible with MNV as it is not possible to
obtain a sufficient quantity of authentic VPg-linked HuNV
Viral protein translation
RNA due to the inability to culture HuNV in immortalized
Once a norovirus VPg-linked RNA genome is released into cells. These studies confirmed that translation of MNV
the cytoplasm of a permissive cell, it behaves as an mRNA RNA requires the eukaryotic initiation factor eIF4A, the
Downloaded from www.microbiologyresearch.org by
280 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

RNA helicase component of eIF4F, potentially for unwind- to the circularization of the viral genome during trans-
ing RNA structures present in the 59 end of the genome, as lation and/or replication. In addition to PCBP2 and
both dominant-negative forms of eIF4A and small- hnRNPA1, RNA interference (RNAi)-mediated reduction
molecule inhibitors of eIF4A inhibited MNV RNA of three of the proteins identified in the proteomics
translation in vitro (Chaudhry et al., 2006). To date only screen, i.e. PTB, DDX3 and La, significantly reduced
direct interactions with the norovirus VPg have been MNV replication in cell culture (Vashist et al., 2012).
demonstrated for eIF3 and eIF4E, but functions for these Importantly, the role of the cellular RNA-binding proteins
interactions have not been described. Despite the direct in norovirus translation is unknown as currently no
interaction, in vitro studies have indicated that eIF4E is not experimental system exists that allows viral RNA trans-
required for the translation of MNV RNA in a rabbit lation in cells to be decoupled from viral RNA replication;
reticulocyte lysate (Chaudhry et al., 2006). Therefore, it is therefore, effects on either process have a negative impact
likely that VPg forms multiple contacts with components on viral replication. It is possible that these types of RNA–
of the translation initiation factor complex, not all of which protein interactions play both positive and negative
are essential for viral RNA translation (Fig. 3). A direct, regulatory roles in the life cycle of noroviruses as data
high-affinity, functional interaction between the norovirus from a related calicivirus indicated that binding of PTB to
VPg and a component of the translation apparatus has yet the viral 59 end impacts negatively virus translation and
to be described. may contribute to the switch from RNA translation to
replication (Karakasiliotis et al., 2010).
The extremities of calicivirus genomes contain evolution-
arily conserved RNA structures that are known to interact Translation of the viral proteins VP1 and VP2 (and VF1 in
with host cell factors to promote viral replication and MNV) occurs primarily from the subgenomic RNA (Fig.
translation (Simmonds et al., 2008; Vashist et al., 2012). 1). This is most likely a strategy to produce higher levels of
For example, structures in the Norwalk virus 59 and 39 ends the major capsid protein for virus assembly as the
are thought to interact with the cellular proteins La, subgenomic RNA is present at higher levels in infected
polypyrimidine tract-binding protein (PTB) and poly(A)- cells than the viral genomic RNA and each icosahedral
binding protein (PABP), and these proteins have also been capsid contains 180 copies of VP1 arranged in 90 dimers
identified as binding RNA structures in the MNV genome (Prasad et al., 1994). Translation of VP2 occurs by a
(Gutiérrez-Escolano et al., 2000b, 2003; Vashist et al., termination–reinitiation mechanism as the norovirus
2012). The function of these interactions has yet to be subgenomic RNA is polycistronic (Napthine et al., 2009).
elucidated fully, but studies with MNV have confirmed Upon termination of ORF2 translation (VP1), ribosomes
that they play a role in virus replication, as is well are thought to remain associated with the RNA to reinitiate
established for other RNA viruses (Bailey et al., 2010b; at the start of ORF3 (VP2). This is facilitated by
Simmonds et al., 2008). Many of the host proteins found to overlapping stop and start codons of ORF2 and ORF3,
interact with norovirus RNA structures have previously respectively (e.g. UAAUG in MNV). An upstream RNA
been identified as regulators of the translation of other motif known as the termination upstream ribosomal-
RNA viruses, e.g. La and PTB stimulate picornavirus binding sequence (TURBS) is also required, which is in
internal ribosome entry site (IRES)-mediated translation part complementary to 18S rRNA and is thought to tether
(reviewed by Fitzgerald & Semler, 2009). Therefore, the ribosome through direct RNA–RNA base pair interac-
binding of these and other factors may serve to enhance tions (Meyers, 2007; Napthine et al., 2009). There is also
viral protein translation, perhaps through stabilization of evidence from transfection of bovine norovirus RNA that
long-range RNA interactions that promote the circulariza- translation of VP1 occurs by a similar termination–
tion of the norovirus genome, as a process of RNA reinitiation between ORF1 and ORF2 on the genomic
circularization is known to stimulate host cell mRNA RNA (McCormick et al., 2008). This was proposed as a
translation (Wells et al., 1998). In both Norwalk virus and unique translation mechanism for GIII viruses and
MNV, the interaction of complementary sequences in the although it is feasible that it could contribute to VP1
59 and 39 extremities of the viral genome is stimulated by translation in other genogroups, the mechanism has yet to
cellular proteins (López-Manrı́quez et al., 2013; Sandoval- be demonstrated within the context of viral infection for
Jaime & Gutiérrez-Escolano, 2009). Two cellular proteins any norovirus. The start codon for translation of ORF4 in
in particular, the poly(C)-binding protein PCBP2 and the MNV is positioned 13 bases downstream of the start of
heterogeneous nuclear ribonucleoprotein hnRNPA1, are ORF2 and may be initiated by leaky scanning and a 22 slip
involved in the circularization of the MNV genome in the ribosome (McFadden et al., 2011).
(López-Manrı́quez et al., 2013). A number of host factors
have also been identified as binding to the extremities of
Replication complex formation
the MNV genome using a proteomics-based approach and
summary tables of all these interactions can be found in Translation of the ORF1 polyprotein is followed by co- and
Vashist et al. (2012). As many of these factors form direct post-translational processing by the viral NS6 protease, and
protein–protein interactions with each other, it is possible results in the release of the viral NS proteins ready for
that numerous cellular RNA-binding proteins contribute replication complex formation and their precursors, some
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 281
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

1. Attachment
factor binding

2. Receptor
binding

VPg
3. Entry AAAA(n)
4. Uncoating

5. Translation

AAAA(n)

6. Polyprotein
VPg RdRp processing 7. Replication complex
formation

8. RNA replication AAAA(n)

de novo initiation
AAAA(n)
10. Exit
VPg-dependent
initiation
AAAA(n)

AAAA(n)

Genomic
RNA
AAAA(n)
AAAA(n)
AAAA(n)
AAAA(n)
AAAA(n)
AAAA(n)
Subgenomic
9. Virion assembly RNA

Fig. 2. Outline of the norovirus life cycle. (1) HuNV and MNV are thought to attach to the cell surface using various carbohydrate
attachment factors. This is not sufficient to mediate entry and binding to an unidentified protein receptor is thought to be
required (2). Entry (3) and uncoating (4) proceed through as-yet-undefined pathways. (5) The incoming viral genome is
translated, through interactions with VPg at the 59 end of the genome (red triangle) and the cellular translation machinery. (6)
The ORF1 polyprotein is co- and post-translationally cleaved by the viral protease NS6. (7) The replication complex is formed by
recruitment of cellular membranes to the perinuclear region of the cell (not shown), through interactions in part with NS1/2 and
NS4. (8) Genome replication occurs via a negative-strand intermediate, and genomic and subgenomic RNA are generated by
the viral RdRp (NS7), using both de novo and VPg-dependent mechanisms of RNA synthesis. (9) The replicated genomes are
translated (within the replication complex) or packaged into the capsid, VP1, for virion assembly and exit (10).

Downloaded from www.microbiologyresearch.org by


282 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

(a) Cap-dependent initiation and was adventitiously solved with the C terminus of an
adjacent NS6 molecule bound in the active site as a peptide
eIF3
PABP
substrate, which facilitated identification of the key
residues involved in peptide binding and specificity (Leen
(n)A
AA et al., 2012).
A
Similar to other positive-strand RNA viruses, norovirus
eIF4G
replication occurs in close association with host-derived
eIF4E eIF4A membrane complexes in the cytoplasm (Belov & van
Kuppeveld, 2012; Wobus et al., 2004) (Fig. 2). Evidence for
m7GTP the replication complex comes from studies with MNV, as
(b) VPg-dependent initiation to date the mechanism of replication complex formation
has not been investigated in cells replicating HuNV RNA.
eIF3 PABP MNV infection induces the formation of membranous
(n)A vesicle clusters in infected cells, and proliferation of the
AA
A viral RNA polymerase (NS7) and dsRNA, a viral RNA
replication intermediate, occurs at punctate foci in the
eIF4G
eIF4A perinuclear region (Hyde et al., 2009; Wobus et al., 2004).
eIF4E The NS proteins, as well as the major and minor capsid
proteins co-localize with NS7 and dsRNA at this location,
VPg forming the replication complex (Hyde & Mackenzie, 2010;
Thorne et al., 2012). More recently the MNV replication
=Biochemically confirmed interactions
complex was shown to be juxtaposed with the microtubule
organizing centre (MTOC) within the perinuclear region,
Fig. 3. Overview of the norovirus translation initiation complex. (a) suggesting that MNV may use the cytoskeletal network
Schematic representation of the eIF4F cap-dependent cellular early in infection to position the complex (Hyde et al.,
translation initiation factor complex formed on the 59 end of capped 2012). This process is mediated potentially by an
cellular mRNAs. The interaction of the eIF4E cap-binding protein interaction between VP1 and acetylated tubulin, although
with the methylated cap structure (m7GTP) recruits the eIF4G a direct interaction in infected cells has yet to be
scaffold. Binding of eIF4G to eIF4E also enables the recruitment of demonstrated.
the RNA helicase component eIF4A and the eIF3 complex. eIF3 is
responsible subsequently for the recruitment of the 40S small The host membranes that comprise the MNV replication
ribosomal subunit prior to translation initiation. (b) Norovirus VPg- complex are derived from components of the secretory
dependent translation also requires an interaction between VPg pathway, notably the endoplasmic reticulum (ER), trans-
and translation initiation factors. Thus far only direct interactions Golgi network and endosomes, which together serve as a
with eIF4E and eIF3 have been documented. Note that the platform for replication (Hyde et al., 2009). The mechan-
interaction of VPg with eIF4E is not required for virus translation isms by which noroviruses recruit cellular membranes to
initiation in vitro. Biochemically established interactions are establish the replication complex in the host cell remain
indicated. unclear. For MNV, NS1/2 and NS4 have been implicated in
driving complex formation by recruiting cellular mem-
branes, based on their localization to components of the
endocytic and secretory pathways (Bailey et al., 2010a;
of which are thought to be functionally active in replication Hyde & Mackenzie, 2010). Furthermore, the MNV NS4
(Belliot et al., 2005). The structure and function of the protein induces Golgi disassembly and mildly inhibits
HuNV protease (Pro, Fig. 1) have been characterized protein secretion, demonstrating that it is capable of
extensively, and the active site is thought to be a catalytic driving the rearrangement of cellular membranes (Sharp
triad of H30, E54 and C139, although E54 is thought et al., 2012). The HuNV NS4 equivalent, p22, inhibits
largely to determine substrate specificity rather than being cellular protein secretion and has a more potent inhibitory
essential for catalysis (Someya et al., 2002, 2008; Zeitler effect on this pathway than MNV NS4 (Sharp et al., 2010,
et al., 2006). HuNV Pro and MNV NS6 share .60 % 2012). HuNV p22 contains an ER export signal that is
amino acid identity (Leen et al., 2012), and sequence thought to promote its uptake into COPII vesicles involved
alignments showed that H30 and C139 are conserved, in transport from the ER to the Golgi apparatus, although a
whereas position 54 is aspartic acid rather than glutamic direct interaction between p22 and COPII has not yet
acid in MNV NS6, although this implies it has a similar been demonstrated. Nevertheless, uptake of p22 has been
function. Taken together with structural and mutagenic proposed to result in mislocalization of COPII vesicles,
studies it is therefore thought that the catalytic triad and thereby inhibiting proper trafficking to the Golgi, and in
mechanism of catalysis may be conserved between HuNV turn leading to Golgi disassembly and loss of protein
3CLPro and MNV NS6 (Leen et al., 2012). The structure of secretion (Sharp et al., 2010). MNV NS4 does not contain
MNV NS6 has recently been determined to high resolution the same ER export signal, and the possible involvement of
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 283
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

COPI and COPII vesicles in MNV replication complex This assay demonstrated that specific loop sequences in the
formation has been excluded (Hyde et al., 2009), suggest- shell domain of VP1 interact with the RdRp to specifically
ing NS4 may act by an alternative mechanism. The HuNV stimulate de novo initiation, without affecting VPg-
p48 protein (equivalent to the MNV NS1/2 protein) dependent initiation. These observations have resulted in
promotes Golgi disassembly, and disrupts expression and a model for the initial rounds of viral RNA synthesis,
trafficking of cell surface proteins in transfected cells whereby the shell domain from the incoming parental viral
(Ettayebi & Hardy, 2003; Fernandez-Vega et al., 2004). capsid, or from the early rounds of viral RNA translation,
Disruption by p48 is thought to be mediated through a binds to the RdRp and stimulates negative-sense RNA
direct interaction with VAP-A (SNARE regulator vesicle- synthesis. The model would propose that as the levels of
associated membrane protein-associated protein A), which VP1 increase due to the production of viral positive-sense
is involved in regulating cellular vesicle transport (Ettayebi RNA, VP1 forms multimeric complexes, preventing the
& Hardy, 2003). Studies on HuNV proteins have focussed interaction with the RdRp and leading to capsid assembly
largely on the disruption to protein secretion; however, it is and the formation of new infectious viral particles. This
plausible that the observed membrane rearrangements model would also explain the observation that bovine
either contribute to or are a direct consequence of the noroviruses have a termination–reinitiation sequence
recruitment of cellular membranes for the replication (TURBS) upstream of VP1, similar to that used to produce
complex. VP2 (McCormick et al., 2008). The effect of this strategy is
that a small percentage of ribosomes translating ORF1 will
Genome replication reinitiate on ORF2 to produce VP1, leading to low levels of
VP1 during the ‘pioneer’ rounds of viral translation.
As for all positive-sense RNA viruses, genome replication Whether other noroviruses utilize a similar mechanism for
occurs via a negative-sense RNA intermediate and is the synthesis of VP1 during the early stages of the viral life
performed by the viral RdRp. The norovirus RdRps, cycle is unknown currently. Species-specific interactions
generally referred to as NS7, contain active-site residues with NS1/2 also promote norovirus RdRp activity in this
conserved among other positive-strand RNA virus RdRps, assay, whereas interactions with VP2 are inhibitory (Subba-
and are conserved structurally and functionally (Högbom Reddy et al., 2011). Taken together, this suggests that the
et al., 2009). The replicative properties of the MNV RdRp viral proteins may play alternative regulatory roles, perhaps
(NS7) are comparable in vitro with those of the HuNV at different stages in the norovirus life cycle. Recent work
RdRp, supporting the use of MNV for studies of has highlighted that the HuNV RdRp may be regulated by
replication (Bull et al., 2011). Subsequent to the ‘pioneer phosphorylation by the cellular survival kinase Akt (Eden et
round’ of translation of the incoming parental viral RNA al., 2011). Akt phosphorylates HuNV NS7 RdRp on residue
genome, the mRNA template then functions as a template T33, which is conserved among pandemic HuNV GII.4
for the formation of a double-stranded replicative form strains, but absent in non-pandemic strains. However, the
(RF). The process of initiation of negative-sense RNA implications of Akt regulation of RdRp activity on infection
synthesis on the incoming parental viral RNA is not have not been investigated fully.
understood fully, but the norovirus RdRp has two
mechanisms of initiation that have been demonstrated Following the generation of a double-stranded RF, the
both in vitro and in cells: de novo and VPg-dependent synthesis of positive-sense genomic and subgenomic RNA
(Rohayem et al., 2006). Unlike for picornavirus replication occurs (Fig. 2). It is accepted generally that this process
where the negative-sense RNA is VPg-linked (Pettersson occurs in a VPg-dependent manner as positive-sense
et al., 1978), there is no evidence as yet to suggest whether infectious norovirus RNA isolated from MNV-infected cells
or not norovirus negative-sense RNA in the double- is linked to VPg (Chaudhry et al., 2006). For this process to
stranded RF is linked to VPg. However, a recently proposed occur, the NS7 RdRp uses VPg as a protein primer for RNA
model has suggested that de novo initiation is used for the synthesis, initiating at the 39 end of the negative-sense RNA.
synthesis of negative-sense genomic and subgenomic RNAs This mechanism is also employed by other caliciviruses and
(Subba-Reddy et al., 2012), both of which can be detected members of the family Picornaviridae of positive-sense RNA
in calicivirus-infected cells (Green et al., 2002). The de novo viruses (Liu et al., 2009; Rohayem et al., 2006). In the case of
initiation of norovirus RdRps is enhanced through direct noroviruses, the NS7 RdRp initiates RNA synthesis by
interactions with the shell domain of VP1, in a species- covalently attaching VPg via a phosphodiester bond to the
specific and concentration-dependent manner (Subba- initiating nucleotide, which is invariably guanine in all
Reddy et al., 2012). This observation was made using members of the family Caliciviridae. This process is referred
cell-based assays that rely on the co-expression of to as VPg nucleotidylation or guanylation. The linkage is
norovirus proteins, including the NS7 RdRp in cells. De formed between the guanine at the 59 end of the genome and
novo RNA synthesis is then measured indirectly via the a conserved tyrosine residue in VPg (Y26 in MNV and Y27
RIG-I-mediated sensing of 59-triphosphorylated RNA in HuNV) (Subba-Reddy et al., 2011). This linkage is
produced by the RdRp using host cell RNAs as a template, essential for infectivity as enzymic removal of VPg from
which subsequently leads to the activation of the IFN-b purified norovirus RNA significantly reduces infectivity
promoter linked to luciferase (Subba-Reddy et al., 2011). (Chaudhry et al., 2006; Guix et al., 2007). Biochemical and
Downloaded from www.microbiologyresearch.org by
284 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

(a) Premature termination (b) Internal initiation

5′(+) 3′ 5′(+) 3′
AAAA(n) AAAA(n)
VPg VPg
Pol/NS7 Pol/NS7
Termination signal
5′(+) 3′ 5′(+) 3′
AAAA(n) AAAA(n)
UUUU(n) UUUU(n)
Termination of 5′(–) 5′(–)
RNA synthesis SG promoter

Internal initiation
on SG promoter

UUUU(n) UUUU(n)
3′ 5′(–) 3′ 5′(–)

5′(+) 3′ 3′
AAAA(n) AAAA(n)
3′ UUUU(n) UUUU(n)
3′
5′(–) 5′(–)

5′(+) 3′
AAAA(n)

5′(+) 3′
AAAA(n)
3′ UUUU(n)
5′(–)

Fig. 4. Mechanisms of norovirus subgenomic RNA synthesis. Schematic representation of the two proposed models for
norovirus subgenomic RNA synthesis. (a) The presence of a termination signal upstream of the VP1-coding region (termination
signal) results in premature termination during negative-sense RNA synthesis by the viral RNA polymerase (NS7/Pol). The VPg-
linked positive-sense RNA template is drawn in black and initiation occurs de novo. The resulting negative-sense subgenomic
RNA (blue) is then used for VPg-dependent RNA synthesis to produce a subgenomic dsRNA. The newly synthesized positive-
sense ‘daughter’ subgenomic RNA (green) can then be used as a template for the synthesis of negative-sense subgenomic
RNA. These in turn function as templates for additional rounds of ‘daughter’ subgenomic RNA synthesis. (b) The internal
initiation model relies on the VPg-dependent subgenomic RNA synthesis occurring on a promoter sequence present
downstream of the VP1-coding region on the negative-sense RNA (blue, SG promoter). The viral RNA polymerase initiates RNA
synthesis, presumably in a VPg-dependent manner, to produce new ‘daughter’ VPg-linked subgenomic RNA (green), which
may in turn function as a template for additional rounds of subgenomic RNA synthesis as shown in the premature termination
model.

structural studies have indicated that the conformation of located in the NS7-coding region (Victoria et al., 2009);
VPg and its helical core is central in facilitating the VPg–RdRp however, this structure is not required for in vitro nucleotidyla-
interaction, but have also highlighted that VPg is likely to tion of the norovirus VPg. There is in fact evidence than an
unfold during the process of guanylation to enable the element at the 39 end of the genome enhances the efficiency of
positioning of the tyrosine in the active site of the RdRp (Leen nucleotidylylation, at least in vitro (Belliot et al., 2005).
et al., 2013). This process of VPg nucleotidylation has been
studied widely in the picornaviruses where a small RNA Based on the observation that both positive- and negative-
structure, first identified in the 2C-coding region of poliovirus sense forms of the viral genomic and subgenomic RNAs
(Goodfellow et al., 2000, 2003) was used as a template to have been detected in calicivirus-infected cells (Green
produce VPg-pUpU that was then used to initiate viral RNA et al., 2002), two models have been put forward for
synthesis at the termini of the template RNAs (Liu et al., the mechanism of norovirus subgenomic RNA synthesis
2009). A similar structure has been proposed for noroviruses, (Fig. 4). The first involves premature termination during
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 285
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

synthesis of the negative-sense genomic RNA, arising from To create a beneficial cellular environment for replication,
a termination signal. The resulting negative-sense sub- noroviruses are thought to disrupt a number of host cell
genomic RNA would then serve as a template for the pathways. As described above, the host cell protein
production of positive-sense subgenomic RNA. The second secretion pathway is one such target that is disrupted
model is that an RNA secondary structure upstream of during replication of both HuNV and MNV to modify and
ORF2 in the negative-sense genomic RNA acts as a recruit membranes for replication (Sharp et al., 2010, 2012;
promoter for synthesis of positive-sense subgenomic Wobus et al., 2004). It is possible, however, that disruption
RNA. In agreement with this model, bioinformatic analysis of this pathway and abrogation of cell surface protein
has indicated the presence of a highly conserved RNA expression confers additional advantages during norovirus
stem–loop structure downstream of the VP1-coding region infections, such as interfering with antigen presentation or
on the negative-sense RNA, 6 nt from the start of the cytokine secretion, as has been reported for other RNA
subgenomic RNA in all members of the family Caliciviridae viruses that disrupt the host cell secretion pathway (Dodd
(Simmonds et al., 2008). Mutational analysis of this et al., 2001). The HuNV replicon system has facilitated a
structure confirmed its essential role in the norovirus life number of studies into HuNV replication in the absence of
cycle, but whether it plays a direct role in viral subgenomic an efficient cell culture system. It consists of a human
RNA synthesis is unknown. In vitro biochemical studies hepatoma cell line (Huh-7) expressing self-replicating
with a related member of the genus Lagovirus, rabbit Norwalk virus genomic RNA and has been used to
haemorrhagic disease virus, suggests that the calicivirus investigate cellular pathways that are altered by HuNV
RdRp can utilize antisense genomic RNA as a template for replication (Chang, 2009; Chang et al., 2006). In this way,
the production of subgenomic-like RNA in vitro, at least both cholesterol and carbohydrate biosynthesis pathways
(Morales et al., 2004). Further studies in this area are were found to be altered in replicon-bearing cells by DNA
warranted to determine if this model is correct, but it is microarray analysis. Cholesterol biosynthesis was down-
worth noting that the models are not mutually exclusive as regulated and the use of statins to lower cholesterol was
positive-sense subgenomic RNA produced in the internal found to promote replication (Chang, 2009). Likewise,
initiation model could be used as a template for replication simvastatin has been found to increase replication in
by NS7. Once the synthesis of positive-sense viral genomic gnotobiotic pigs infected with HuNV (Jung et al., 2012)
and subgenomic RNAs is initiated, multiple rounds of and epidemiological data has since raised concerns
translation of the newly synthesized RNAs ensues, followed regarding the use of statins as an increased risk factor
by additional rounds of RNA synthesis. in the elderly (Rondy et al., 2011). The enhanced
replication resulting from treatment with statins in vivo is
thought in part to be mediated through suppression of
Interactions with host cell factors and pathways innate immunity (Jung et al., 2012); however, the direct
molecular basis for the interaction between norovirus and
Interactions with host cell proteins and cellular pathways cholesterol synthesis has yet to be determined.
are essential for achieving the productive replication of all
positive-strand RNA viruses (Nagy & Pogany, 2011). Studies on related positive-sense RNA viruses suggest an
Knowledge of host cell interactions involved in the extensive modification of the cellular transcriptional and
norovirus life cycle currently lags behind that of other translational apparatus occurs with infection, which is
RNA viruses. As mentioned above, several studies have beneficial to virus replication. In addition to the modification
demonstrated interactions between host cell proteins and of the cholesterol pathway, in vitro biochemical studies have
structures in the viral RNA, which play a role in the viral life indicated that the norovirus NS6 protease possess the ability
cycle (Bailey et al., 2010b; Gutiérrez-Escolano et al., 2000b; to cleave components of the cellular translation apparatus,
Sandoval-Jaime & Gutiérrez-Escolano, 2009; Vashist et al., namely PABP (Kuyumcu-Martinez et al., 2004). Cleavage of
2012). The host cell proteins DDX3, La and PTB were all PABP, as well as other components of the eIF4F complex, has
found subsequently to at least partially localize to the viral been observed during FCV infection (Willcocks et al., 2004),
replication complex during infection, and RNAi-mediated but whether or not this occurs during authentic norovirus
knockdown of these factors reduced MNV replication in replication has yet to be determined.
cell culture, indicating that these interactions play a role in Modification of the nuclear–cytoplasmic shuttling path-
replication, although their specific functions are not known ways has also been reported in other positive-sense RNA
(Vashist et al., 2012). Another study identified an viruses and several observations would suggest that this
interaction between PTB and the polypyrimidine tract in may also occur during norovirus infection of permissive
a stem–loop structure at the 39 end of the MNV genome cells. (i) Infection with FCV results in the movement of the
(Bailey et al., 2010b). Mutation of the polypyrimidine tract PTB protein from the nucleus to the cytoplasm where it
reduced PTB binding and had no effect on replication in negatively regulates FCV translation (Karakasiliotis et al.,
cell culture, but was sufficient to attenuate MNV in vivo, 2010). Importantly, this appears to occur prior to a global
demonstrating a role of RNA–protein interactions in effect on nuclear export. (ii) During MNV infection of cells
norovirus replication and as possible determinants of the nuclear–cytoplasmic shuttling protein hnRNPA1,
pathogenesis (Bailey et al., 2010b). predominantly nuclear in uninfected macrophage cells,
Downloaded from www.microbiologyresearch.org by
286 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

redistributes to the cytoplasm upon infection, where, as assembly during virus replication (Bertolotti-Ciarlet et al.,
highlighted above, it is thought to play a role in norovirus 2002), although the involvement of cellular proteins cannot
genome circularization (López-Manrı́quez et al., 2013). be excluded. This self-assembly property of the norovirus
VP1 protein has been utilized to generate a recombinant
Antagonistic interactions with the innate immune system are
VLP-based vaccine that has proven efficacious in human
a well-known feature of RNA virus infections. Given the short
volunteer studies (Atmar et al., 2011). Whilst VP2 is not
time course of HuNV, innate immunity has been implicated
required for VLP assembly, it is thought to promote the
highly in the control of infection, which has been supported
stability of VP1 and is essential for the production of
by studies with HuNV and MNV. Treatment with IFN-a/b
infectious virions (Bertolotti-Ciarlet et al., 2002; Sosnovtsev
reduced HuNV RNA replication in the replicon system in
et al., 2005). The highly basic nature of VP2 has formed the
gnotobiotic pigs and decreased replication of MNV in vitro
basis of a long-held theory that it may be involved in
(Chang & George, 2007; Jung et al., 2012). The role of the
encapsidation via an interaction with the acidic viral RNA
signal transducer and activator of transcription STAT1
(Sosnovtsev et al., 2005). In support of this, VP1 and VP2
signalling in the IFN-mediated innate immune response has
have been shown to interact, and the binding of VP2 was
also been well documented to control the appearance of
mapped to a conserved motif in the shell domain of VP1
clinical disease and viral dissemination during MNV infection
(Vongpunsawad et al., 2013). This places VP2 in the interior
in vivo (Karst et al., 2003; Mumphrey et al., 2007; Wobus et al.,
of the capsid, consistent with a role in encapsidation.
2004). The initial detection of norovirus RNA is thought to
However, to date, a direct interaction between VP2 and viral
occur primarily by the cellular helicase MDA-5, which senses
genomic RNA has not been demonstrated. A possible
dsRNA, rather than the RIG-I RNA sensor, which did not
interaction between the VPg protein linked covalently to the
confer resistance to HuNV replication in transfected Huh-7
viral RNA and the capsid protein VP1 has also been
cells (Guix et al., 2007; McCartney et al., 2008). Accordingly,
observed in yeast two-hybrid studies with the related FCV
MNV replicates to higher titres in MDA-52/2 mice and also,
(Kaiser et al., 2006). Such an interaction would provide a
to a smaller extent, higher in TLR32/2 mice, suggesting it may
mechanism for the specific encapsidation of replicated viral
also contribute to detection of dsRNA (McCartney et al.,
RNA in preference to cellular mRNAs, although further
2008). The signalling cascade initiated by MDA-5 is mediated
studies are required to validate such a model.
by the mitochondrial antiviral signalling (MAVS) protein,
which results in activation of the transcription factors IFN- There have been no direct studies on the exit and release of
regulatory factor (IRF)-3, IRF-7 and NF-kB. This in turn the assembled norovirus virion to complete the viral life cycle.
results in the production of type I IFNs and the upregulation One proposed exit strategy for other caliciviruses involves the
of IFN-stimulated genes (ISGs) (Kawai et al., 2005; Seth et al., induction of apoptosis (Alonso et al., 1998; Bok et al., 2009;
2005). Both IRF-3 and IRF-7 are required for IFN-mediated Sosnovtsev et al., 2003). Whether noroviruses employ a
control of MNV (Thackray et al., 2012). MNV is able to similar strategy is unknown; however, accumulation of
interfere with innate immune signalling at the cellular level at apoptotic epithelial cells has been observed in intestinal
least partially through the actions of VF1, the product of biopsies from infected patients (Bok et al., 2009; Furman et al.,
ORF4 (McFadden et al., 2011). In vitro, VF1 was found 2009; Troeger et al., 2009). Furthermore, as highlighted
recently to delay the upregulation of IFN-b and other ISGs, above, apoptosis is induced with active MNV replication and
and loss of VF1 resulted in a fitness cost in vitro and plays a role in the processing of NS1/2 during the latter stages
attenuation in vivo (McFadden et al., 2011). Interestingly of the viral life cycle. The induction is associated with
attenuation was observed in a STAT12/2 model of infection, downregulation of survivin, a pro-survival factor, followed by
suggesting that STAT1-independent signalling pathways may caspase and cathepsin B activation and cytochrome c release
also be involved in, but are not sufficient to control, MNV (Bok et al., 2009; Furman et al., 2009). Whether the
infection alone. Given its mitochondrial localization, VF1 is downregulation of survivin occurs by a direct interaction
expected to interfere with signalling through the MAVS with a viral protein or as a response to indirect effects of MNV
complex or a downstream component affecting IRF-3 and on the host cell is as yet unknown. Instead, VF1 has been
IRF-7 activation, although further studies are required to shown to delay the onset of apoptosis through its localization
elucidate the direct target. As HuNV does not express a at the mitochondria (McFadden et al., 2011), although it is
homologue to VF1, it is unknown at present if it has the same possible that this may serve to prolong the window for
capacity as MNV to antagonize the innate immune response; replication prior to the final stages of exit and release.
however, functional duplication of VF1 remains a possibility. Inhibition of apoptosis results in reduced MNV production,
either way demonstrating a clear requirement for apoptosis
for an aspect of the norovirus life cycle (Furman et al., 2009).
Assembly and exit
The processes behind viral assembly, encapsidation and the
Concluding remarks
exit of noroviruses are largely unknown. The ability of VP1
to self-assemble into virus-like particles (VLPs) indistin- Although our understanding of the norovirus life cycle is
guishable morphologically and antigenically from native far from complete, the past decade has seen a dramatic
virions suggests that it may be sufficient to drive capsid increase in publications relating to various aspects of
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 287
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

norovirus biology, including the molecular mechanisms of Bailey, D., Kaiser, W. J., Hollinshead, M., Moffat, K., Chaudhry, Y.,
norovirus genome translation and replication. This is in Wileman, T., Sosnovtsev, S. V. & Goodfellow, I. G. (2010a). Feline
calicivirus p32, p39 and p30 proteins localize to the endoplasmic
part due to increased awareness of the prevalence and
reticulum to initiate replication complex formation. J Gen Virol 91,
clinical importance of noroviruses, but also due to the 739–749.
significant efforts in the development of tools and
Bailey, D., Karakasiliotis, I., Vashist, S., Chung, L. M. W., Rees, J.,
resources for the study of norovirus biology. A lack of an McFadden, N., Benson, A., Yarovinsky, F., Simmonds, P. & Goodfellow,
efficient culture system for HuNVs is a major hurdle that I. (2010b). Functional analysis of RNA structures present at the 39
researchers in the field have yet to overcome, but steps extremity of the murine norovirus genome: the variable polypyrimidine
towards this are already under way (Asanaka et al., 2005; tract plays a role in viral virulence. J Virol 84, 2859–2870.
Guix et al., 2007). Work with related viruses has provided Belliot, G., Sosnovtsev, S. V., Chang, K.-O., Babu, V., Uche, U.,
novel insights into the biology of members of the family Arnold, J. J., Cameron, C. E. & Green, K. Y. (2005). Norovirus
Caliciviridae, but mechanistic differences between the proteinase-polymerase and polymerase are both active forms of RNA-
various members of the family are apparent. For example, dependent RNA polymerase. J Virol 79, 2393–2403.
whilst further processing of the MNV NS1/2 protein occurs Belov, G. A. & van Kuppeveld, F. J. M. (2012). (+)RNA viruses rewire
during virus replication, whether the HuNV p48 (N-term) cellular pathways to build replication organelles. Curr Opin Virol 2,
740–747.
protein is further processed in not known. Also, whilst FCV
translation initiation is clearly dependent on eIF4E, the Bertolotti-Ciarlet, A., White, L. J., Chen, R., Prasad, B. V. V. & Estes,
M. K. (2002). Structural requirements for the assembly of Norwalk
only study to date on norovirus translation initiation virus-like particles. J Virol 76, 4044–4055.
found that eIF4E is in fact dispensable (Chaudhry et al.,
Bok, K. & Green, K. Y. (2012). Norovirus gastroenteritis in
2006). With this in mind, continued efforts and the use of immunocompromised patients. N Engl J Med 367, 2126–2132.
novel approaches to norovirus culture will be required to
Bok, K., Prikhodko, V. G., Green, K. Y. & Sosnovtsev, S. V. (2009).
provide the breakthrough needed. One such approach has Apoptosis in murine norovirus-infected RAW264.7 cells is associated
found recently that mice deficient in aspects of the immune with downregulation of survivin. J Virol 83, 3647–3656.
response can be infected by intraperitoneal injection of
Bull, R. A., Hyde, J., Mackenzie, J. M., Hansman, G. S., Oka, T.,
HuNV (Taube et al., 2013). This experimental system, which Takeda, N. & White, P. A. (2011). Comparison of the replication
does not recapitulate norovirus gastrointestinal disease, properties of murine and human calicivirus RNA-dependent RNA
does, however, provide a useful tool for the study of aspects polymerases. Virus Genes 42, 16–27.
of norovirus pathogenesis. More importantly, this system, Centers for Disease Control and Prevention (CDC) (2002).
combined with the advances in other areas, provides a Outbreak of acute gastroenteritis associated with Norwalk-like viruses
starting block from which to move forward to the ‘Holy among British military personnel – Afghanistan, May 2002. MMWR
Grail’ of norovirus biology, i.e. a cell culture system that Morb Mortal Wkly Rep 51, 477–479.
enables HuNVs to undergo a full replication cycle. Such a Chan, W. K. Y., Lee, K. W. & Fan, T. W. (2010). Pneumatosis intestinalis
system, when combined with a reverse-genetics system to in a child with nephrotic syndrome and norovirus gastroenteritis.
Pediatr Nephrol 25, 1563–1566.
enable viral genome manipulation and a small, preferably
inexpensive and genetically defined, animal model, would Chang, K.-O. (2009). Role of cholesterol pathways in norovirus
replication. J Virol 83, 8587–8595.
provide the complete toolset with which to fill in the
remaining gaps in our understanding of norovirus biology. Chang, K.-O. & George, D. W. (2007). Interferons and ribavirin
effectively inhibit Norwalk virus replication in replicon-bearing cells.
J Virol 81, 12111–12118.
Acknowledgements Chang, K.-O., Sosnovtsev, S. V., Belliot, G., King, A. D. & Green, K. Y.
(2006). Stable expression of a Norwalk virus RNA replicon in a
The authors are supported by funding from the Wellcome Trust as a human hepatoma cell line. Virology 353, 463–473.
PhD Studentship to L. G. T. and a Senior Fellowship to I. G. G. I. G. G.
Chaudhry, Y., Nayak, A., Bordeleau, M.-E., Tanaka, J., Pelletier, J.,
is also supported by funding from the Biotechnology and Biological
Belsham, G. J., Roberts, L. O. & Goodfellow, I. G. (2006).
Sciences Research Council.
Caliciviruses differ in their functional requirements for eIF4F
components. J Biol Chem 281, 25315–25325.
References Chaudhry, Y., Skinner, M. A. & Goodfellow, I. G. (2007). Recovery of
genetically defined murine norovirus in tissue culture by using a
Alonso, C., Oviedo, J. M., Martı́n-Alonso, J. M., Dı́az, E., Boga, J. A. & fowlpox virus expressing T7 RNA polymerase. J Gen Virol 88, 2091–
Parra, F. (1998). Programmed cell death in the pathogenesis of rabbit 2100.
hemorrhagic disease. Arch Virol 143, 321–332. Cheetham, S., Souza, M., Meulia, T., Grimes, S., Han, M. G. & Saif,
Asanaka, M., Atmar, R. L., Ruvolo, V., Crawford, S. E., Neill, F. H. & L. J. (2006). Pathogenesis of a genogroup II human norovirus in
Estes, M. K. (2005). Replication and packaging of Norwalk virus RNA gnotobiotic pigs. J Virol 80, 10372–10381.
in cultured mammalian cells. Proc Natl Acad Sci U S A 102, 10327– Clarke, I. N. & Lambden, P. R. (2000). Organization and expression of
10332. calicivirus genes. J Infect Dis 181 (Suppl 2), S309–S316.
Atmar, R. L., Bernstein, D. I., Harro, C. D., Al-Ibrahim, M. S., Chen, Daughenbaugh, K. F., Fraser, C. S., Hershey, J. W. B. & Hardy, M. E.
W. H., Ferreira, J., Estes, M. K., Graham, D. Y., Opekun, A. R. & other (2003). The genome-linked protein VPg of the Norwalk virus binds
authors (2011). Norovirus vaccine against experimental human eIF3, suggesting its role in translation initiation complex recruitment.
Norwalk virus illness. N Engl J Med 365, 2178–2187. EMBO J 22, 2852–2859.
Downloaded from www.microbiologyresearch.org by
288 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

Dodd, D. A., Giddings, T. H., Jr & Kirkegaard, K. (2001). Poliovirus 3A Guix, S., Asanaka, M., Katayama, K., Crawford, S. E., Neill, F. H.,
protein limits interleukin-6 (IL-6), IL-8, and beta interferon secretion Atmar, R. L. & Estes, M. K. (2007). Norwalk virus RNA is infectious in
during viral infection. J Virol 75, 8158–8165. mammalian cells. J Virol 81, 12238–12248.
Donaldson, E. F., Lindesmith, L. C., Lobue, A. D. & Baric, R. S. (2008). Gutiérrez-Escolano, A. L., Brito, Z. U., del Angel, R. M. & Jiang,
Norovirus pathogenesis: mechanisms of persistence and immune X. (2000a). Interaction of cellular proteins with the 59 end of Norwalk
evasion in human populations. Immunol Rev 225, 190–211. virus genomic RNA. J Virol 74, 8558–8562.
Donaldson, E. F., Lindesmith, L. C., Lobue, A. D. & Baric, R. S. (2010). Gutiérrez-Escolano, A. L., Brito, Z. U., del Angel, R. M. & Jiang,
Viral shape-shifting: norovirus evasion of the human immune system. X. (2000b). Interaction of cellular proteins with the 59 end of Norwalk
Nat Rev Microbiol 8, 231–241. virus genomic RNA. J Virol 74, 8558–8562.
Duizer, E., Schwab, K. J., Neill, F. H., Atmar, R. L., Koopmans, M. P. G. Gutiérrez-Escolano, A. L., Vázquez-Ochoa, M., Escobar-Herrera,
& Estes, M. K. (2004). Laboratory efforts to cultivate noroviruses. J. & Hernández-Acosta, J. (2003). La, PTB, and PAB proteins bind to
J Gen Virol 85, 79–87. the 39 untranslated region of Norwalk virus genomic RNA. Biochem
Biophys Res Commun 311, 759–766.
Eden, J.-S., Sharpe, L. J., White, P. A. & Brown, A. J. (2011). Norovirus
RNA-dependent RNA polymerase is phosphorylated by an important Hall, A. J., Lopman, B. A., Payne, D. C., Patel, M. M., Gastañaduy,
survival kinase, Akt. J Virol 85, 10894–10898. P. A., Vinjé, J. & Parashar, U. D. (2013). Norovirus disease in the
United States. Emerg Infect Dis 19, 1198–1205.
Ettayebi, K. & Hardy, M. E. (2003). Norwalk virus nonstructural
protein p48 forms a complex with the SNARE regulator VAP-A and Herbert, T. P., Brierley, I. & Brown, T. D. (1997). Identification of a
prevents cell surface expression of vesicular stomatitis virus G protein. protein linked to the genomic and subgenomic mRNAs of feline
J Virol 77, 11790–11797. calicivirus and its role in translation. J Gen Virol 78, 1033–1040.
Farkas, T., Nakajima, S., Sugieda, M., Deng, X., Zhong, W. & Jiang, Högbom, M., Jäger, K., Robel, I., Unge, T. & Rohayem, J. (2009).
X. (2005). Seroprevalence of noroviruses in swine. J Clin Microbiol 43, The active form of the norovirus RNA-dependent RNA polymerase
657–661. is a homodimer with cooperative activity. J Gen Virol 90, 281–
291.
Fernandez-Vega, V., Sosnovtsev, S. V., Belliot, G., King, A. D., Mitra,
T., Gorbalenya, A. & Green, K. Y. (2004). Norwalk virus N-terminal Hyde, J. L. & Mackenzie, J. M. (2010). Subcellular localization of the
nonstructural protein is associated with disassembly of the Golgi MNV-1 ORF1 proteins and their potential roles in the formation of
complex in transfected cells. J Virol 78, 4827–4837. the MNV-1 replication complex. Virology 406, 138–148.
Firth, A. E. & Brierley, I. (2012). Non-canonical translation in RNA Hyde, J. L., Sosnovtsev, S. V., Green, K. Y., Wobus, C., Virgin, H. W. &
viruses. J Gen Virol 93, 1385–1409. Mackenzie, J. M. (2009). Mouse norovirus replication is associated
with virus-induced vesicle clusters originating from membranes
Fitzgerald, K. D. & Semler, B. L. (2009). Bridging IRES elements in derived from the secretory pathway. J Virol 83, 9709–9719.
mRNAs to the eukaryotic translation apparatus. Biochim Biophys Acta
1789, 518–528. Hyde, J. L., Gillespie, L. K. & Mackenzie, J. M. (2012). Mouse
norovirus 1 utilizes the cytoskeleton network to establish localization
Fuentes, C., Bosch, A., Pintó, R. M. & Guix, S. (2012). Identification of the replication complex proximal to the microtubule organizing
of human astrovirus genome-linked protein (VPg) essential for virus center. J Virol 86, 4110–4122.
infectivity. J Virol 86, 10070–10078.
Ito, S., Takeshita, S., Nezu, A., Aihara, Y., Usuku, S., Noguchi, Y. &
Furman, L. M., Maaty, W. S., Petersen, L. K., Ettayebi, K., Hardy, M. E. Yokota, S. (2006). Norovirus-associated encephalopathy. Pediatr
& Bothner, B. (2009). Cysteine protease activation and apoptosis in Infect Dis J 25, 651–652.
Murine norovirus infection. Virol J 6, 139.
Jung, K., Wang, Q., Kim, Y., Scheuer, K., Zhang, Z., Shen, Q., Chang,
Gerondopoulos, A., Jackson, T., Monaghan, P., Doyle, N. & Roberts, K.-O. & Saif, L. J. (2012). The effects of simvastatin or interferon-a on
L. O. (2010). Murine norovirus-1 cell entry is mediated through a infectivity of human norovirus using a gnotobiotic pig model for the
non-clathrin-, non-caveolae-, dynamin- and cholesterol-dependent study of antivirals. PLoS ONE 7, e41619.
pathway. J Gen Virol 91, 1428–1438.
Kaiser, W. J., Chaudhry, Y., Sosnovtsev, S. V. & Goodfellow, I. G.
Glass, R. I., Parashar, U. D. & Estes, M. K. (2009). Norovirus (2006). Analysis of protein–protein interactions in the feline
gastroenteritis. N Engl J Med 361, 1776–1785. calicivirus replication complex. J Gen Virol 87, 363–368.
Goodfellow, I. (2011). The genome-linked protein VPg of vertebrate Kapikian, A. Z. (2000). The discovery of the 27-nm Norwalk virus: an
viruses – a multifaceted protein. Curr Opin Virol 1, 355–362. historic perspective. J Infect Dis 181 (Suppl 2), S295–S302.
Goodfellow, I., Chaudhry, Y., Richardson, A., Meredith, J., Almond, Karakasiliotis, I., Vashist, S., Bailey, D., Abente, E. J., Green, K. Y.,
J. W., Barclay, W. & Evans, D. J. (2000). Identification of a cis-acting Roberts, L. O., Sosnovtsev, S. V. & Goodfellow, I. G. (2010).
replication element within the poliovirus coding region. J Virol 74, Polypyrimidine tract binding protein functions as a negative regulator
4590–4600. of feline calicivirus translation. PLoS ONE 5, e9562.
Goodfellow, I. G., Kerrigan, D. & Evans, D. J. (2003). Structure and Karst, S. M., Wobus, C. E., Lay, M., Davidson, J. & Virgin, H. W., IV
function analysis of the poliovirus cis-acting replication element (2003). STAT1-dependent innate immunity to a Norwalk-like virus.
(CRE). RNA 9, 124–137. Science 299, 1575–1578.
Goodfellow, I., Chaudhry, Y., Gioldasi, I., Gerondopoulos, A., Natoni, Kawai, T., Takahashi, K., Sato, S., Coban, C., Kumar, H., Kato, H.,
A., Labrie, L., Laliberté, J.-F. & Roberts, L. (2005). Calicivirus Ishii, K. J., Takeuchi, O. & Akira, S. (2005). IPS-1, an adaptor
translation initiation requires an interaction between VPg and eIF4E. triggering RIG-I- and Mda5-mediated type I interferon induction.
EMBO Rep 6, 968–972. Nat Immunol 6, 981–988.
Green, K. Y., Mory, A., Fogg, M. H., Weisberg, A., Belliot, G., Wagner, Kim, M. J., Kim, Y.-J., Lee, J. H., Lee, J. S., Kim, J. H., Cheon, D. S.,
M., Mitra, T., Ehrenfeld, E., Cameron, C. E. & Sosnovtsev, S. V. Jeong, H. S., Koo, H. H., Sung, K. W. & other authors (2011).
(2002). Isolation of enzymatically active replication complexes from Norovirus: a possible cause of pneumatosis intestinalis. J Pediatr
feline calicivirus-infected cells. J Virol 76, 8582–8595. Gastroenterol Nutr 52, 314–318.
Downloaded from www.microbiologyresearch.org by
http://vir.sgmjournals.org 289
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
L. G. Thorne and I. G Goodfellow

Kuyumcu-Martinez, M., Belliot, G., Sosnovtsev, S. V., Chang, K.-O., Napthine, S., Lever, R. A., Powell, M. L., Jackson, R. J., Brown, T. D. K.
Green, K. Y. & Lloyd, R. E. (2004). Calicivirus 3C-like proteinase & Brierley, I. (2009). Expression of the VP2 protein of murine
inhibits cellular translation by cleavage of poly(A)-binding protein. norovirus by a translation termination-reinitiation strategy. PLoS
J Virol 78, 8172–8182. ONE 4, e8390.
Leen, E. N., Baeza, G. & Curry, S. (2012). Structure of a murine Papafragkou, E., Hewitt, J., Park, G. W., Greening, G. & Vinjé,
norovirus NS6 protease-product complex revealed by adventitious J. (2013). Challenges of culturing human norovirus in three-
crystallisation. PLoS ONE 7, e38723. dimensional organoid intestinal cell culture models. PLoS ONE 8,
Leen, E. N., Kwok, K. Y. R., Birtley, J. R., Simpson, P. J., Subba-
e63485.
Reddy, C. V., Chaudhry, Y., Sosnovtsev, S. V., Green, K. Y., Prater, Patel, M. M., Widdowson, M.-A., Glass, R. I., Akazawa, K., Vinjé, J. &
S. N. & other authors (2013). Structures of the compact helical core Parashar, U. D. (2008). Systematic literature review of role of
domains of feline calicivirus and murine norovirus VPg proteins. noroviruses in sporadic gastroenteritis. Emerg Infect Dis 14, 1224–1231.
J Virol 87, 5318–5330. Perry, J. W. & Wobus, C. E. (2010). Endocytosis of murine norovirus 1
Léonard, S., Plante, D., Wittmann, S., Daigneault, N., Fortin, M. G. & into murine macrophages is dependent on dynamin II and
Laliberté, J. F. (2000). Complex formation between potyvirus VPg cholesterol. J Virol 84, 6163–6176.
and translation eukaryotic initiation factor 4E correlates with virus Pettersson, R. F., Ambros, V. & Baltimore, D. (1978). Identification of
infectivity. J Virol 74, 7730–7737. a protein linked to nascent poliovirus RNA and to the polyuridylic
Liu, Y., Wimmer, E. & Paul, A. V. (2009). Cis-acting RNA elements in acid of negative-strand RNA. J Virol 27, 357–365.
human and animal plus-strand RNA viruses. Biochim Biophys Acta Pletneva, M. A., Sosnovtsev, S. V. & Green, K. Y. (2001). The genome
1789, 495–517. of hawaii virus and its relationship with other members of the
López-Manrı́quez, E., Vashist, S., Ureña, L., Goodfellow, I., Chavez, caliciviridae. Virus Genes 23, 5–16.
P., Mora-Heredia, J. E., Cancio-Lonches, C., Garrido, E. & Gutiérrez- Prasad, B. V., Rothnagel, R., Jiang, X. & Estes, M. K. (1994). Three-
Escolano, A. L. (2013). Norovirus genome circularization and dimensional structure of baculovirus-expressed Norwalk virus
efficient replication are facilitated by binding of PCBP2 and hnRNP capsids. J Virol 68, 5117–5125.
A1. J Virol 87, 11371–11387. Rohayem, J., Robel, I., Jäger, K., Scheffler, U. & Rudolph, W. (2006).
McCartney, S. A., Thackray, L. B., Gitlin, L., Gilfillan, S., Virgin, H. W. & Protein-primed and de novo initiation of RNA synthesis by norovirus
Colonna, M. (2008). MDA-5 recognition of a murine norovirus. PLoS 3Dpol. J Virol 80, 7060–7069.
Pathog 4, e1000108. Rondy, M., Koopmans, M., Rotsaert, C., Van Loon, T., Beljaars, B.,
McCormick, C. J., Salim, O., Lambden, P. R. & Clarke, I. N. (2008). Van Dijk, G., Siebenga, J., Svraka, S., Rossen, J. W. A. & other
Translation termination reinitiation between open reading frame 1 authors (2011). Norovirus disease associated with excess mortality
(ORF1) and ORF2 enables capsid expression in a bovine norovirus and use of statins: a retrospective cohort study of an outbreak
without the need for production of viral subgenomic RNA. J Virol 82, following a pilgrimage to Lourdes. Epidemiol Infect 139, 453–463.
8917–8921. Sandoval-Jaime, C. & Gutiérrez-Escolano, A. L. (2009). Cellular
McFadden, N., Bailey, D., Carrara, G., Benson, A., Chaudhry, Y., proteins mediate 59–39 end contacts of Norwalk virus genomic RNA.
Shortland, A., Heeney, J., Yarovinsky, F., Simmonds, P. & other Virology 387, 322–330.
authors (2011). Norovirus regulation of the innate immune response Seth, R. B., Sun, L., Ea, C.-K. & Chen, Z. J. (2005). Identification and
occurs via the product of the alternative open reading frame 4. PLoS characterization of MAVS, a mitochondrial antiviral signaling protein
Pathog 7, e1002413. that activates NF-kappaB and IRF 3. Cell 122, 669–682.
Medici, M. C., Abelli, L. A., Dodi, I., Dettori, G. & Chezzi, C. (2010). Sharp, T. M., Guix, S., Katayama, K., Crawford, S. E. & Estes, M. K.
Norovirus RNA in the blood of a child with gastroenteritis and (2010). Inhibition of cellular protein secretion by norwalk virus
convulsions – a case report. J Clin Virol 48, 147–149. nonstructural protein p22 requires a mimic of an endoplasmic
Mesquita, J. R., Barclay, L., Nascimento, M. S. J. & Vinjé, J. (2010). reticulum export signal. PLoS ONE 5, e13130.
Novel norovirus in dogs with diarrhea. Emerg Infect Dis 16, 980–982. Sharp, T. M., Crawford, S. E., Ajami, N. J., Neill, F. H., Atmar, R. L.,
Mesquita, J. R., Costantini, V. P., Cannon, J. L., Lin, S.-C., Katayama, K., Utama, B. & Estes, M. K. (2012). Secretory pathway
Nascimento, M. S. J. & Vinjé, J. (2013). Presence of antibodies antagonism by calicivirus homologues of Norwalk virus nonstructural
against genogroup VI norovirus in humans. Virol J 10, 176. protein p22 is restricted to noroviruses. Virol J 9, 181.
Meyers, G. (2007). Characterization of the sequence element directing Simmonds, P., Karakasiliotis, I., Bailey, D., Chaudhry, Y., Evans, D. J.
translation reinitiation in RNA of the calicivirus rabbit hemorrhagic & Goodfellow, I. G. (2008). Bioinformatic and functional analysis of
disease virus. J Virol 81, 9623–9632. RNA secondary structure elements among different genera of human
and animal caliciviruses. Nucleic Acids Res 36, 2530–2546.
Morales, M., Bárcena, J., Ramı́rez, M. A., Boga, J. A., Parra, F. &
Torres, J. M. (2004). Synthesis in vitro of rabbit hemorrhagic disease Someya, Y., Takeda, N. & Miyamura, T. (2002). Identification of
virus subgenomic RNA by internal initiation on (–)sense genomic active-site amino acid residues in the Chiba virus 3C-like protease.
RNA: mapping of a subgenomic promoter. J Biol Chem 279, 17013– J Virol 76, 5949–5958.
17018. Someya, Y., Takeda, N. & Wakita, T. (2008). Saturation mutagenesis
Mumphrey, S. M., Changotra, H., Moore, T. N., Heimann-Nichols, reveals that GLU54 of norovirus 3C-like protease is not essential for
E. R., Wobus, C. E., Reilly, M. J., Moghadamfalahi, M., Shukla, D. & the proteolytic activity. J Biochem 144, 771–780.
Karst, S. M. (2007). Murine norovirus 1 infection is associated with Sosnovtsev, S. & Green, K. Y. (1995). RNA transcripts derived from a
histopathological changes in immunocompetent hosts, but clinical cloned full-length copy of the feline calicivirus genome do not require
disease is prevented by STAT1-dependent interferon responses. J Virol VpG for infectivity. Virology 210, 383–390.
81, 3251–3263. Sosnovtsev, S. V., Prikhod’ko, E. A., Belliot, G., Cohen, J. I. & Green,
Nagy, P. D. & Pogany, J. (2011). The dependence of viral RNA K. Y. (2003). Feline calicivirus replication induces apoptosis in
replication on co-opted host factors. Nat Rev Microbiol 10, 137–149. cultured cells. Virus Res 94, 1–10.
Downloaded from www.microbiologyresearch.org by
290 Journal of General Virology 95
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28
Norovirus gene expression and replication

Sosnovtsev, S. V., Belliot, G., Chang, K.-O., Onwudiwe, O. & Green, functional changes of the duodenum in human norovirus infection.
K. Y. (2005). Feline calicivirus VP2 is essential for the production of Gut 58, 1070–1077.
infectious virions. J Virol 79, 4012–4024. Turcios-Ruiz, R. M., Axelrod, P., St John, K., Bullitt, E., Donahue, J.,
Sosnovtsev, S. V., Belliot, G., Chang, K.-O., Prikhodko, V. G., Robinson, N. & Friss, H. E. (2008). Outbreak of necrotizing
Thackray, L. B., Wobus, C. E., Karst, S. M., Virgin, H. W. & Green, K. Y. enterocolitis caused by norovirus in a neonatal intensive care unit.
(2006). Cleavage map and proteolytic processing of the murine J Pediatr 153, 339–344.
norovirus nonstructural polyprotein in infected cells. J Virol 80,
Vashist, S., Bailey, D., Putics, A. & Goodfellow, I. (2009). Model
7816–7831.
systems for the study of human norovirus biology. Future Virol 4,
Subba-Reddy, C. V., Goodfellow, I. & Kao, C. C. (2011). VPg-primed 353–367.
RNA synthesis of norovirus RNA-dependent RNA polymerases by
Vashist, S., Ureña, L., Chaudhry, Y. & Goodfellow, I. (2012).
using a novel cell-based assay. J Virol 85, 13027–13037.
Identification of RNA–protein interaction networks involved in the
Subba-Reddy, C. V., Yunus, M. A., Goodfellow, I. G. & Kao, C. C. norovirus life cycle. J Virol 86, 11977–11990.
(2012). Norovirus RNA synthesis is modulated by an interaction
Victoria, M., Colina, R., Miagostovich, M. P., Leite, J. P. & Cristina, J.
between the viral RNA-dependent RNA polymerase and the major
(2009). Phylogenetic prediction of cis-acting elements: a cre-like
capsid protein, VP1. J Virol 86, 10138–10149.
sequence in norovirus genome? BMC Res Notes 2, 176.
Takanashi, S., Saif, L. J., Hughes, J. H., Meulia, T., Jung, K., Scheuer, K. A.
& Wang, Q. (2013). Failure of propagation of human norovirus in intestinal Vongpunsawad, S., Venkataram Prasad, B. V. & Estes, M. K. (2013).
epithelial cells with microvilli grown in three-dimensional cultures. Arch Norwalk virus minor capsid protein VP2 associates within the VP1
Virol. doi: 10.1007/s00705-013-1806-4 [Epub ahead of print]. shell domain. J Virol 87, 4818–4825.
Taube, S., Perry, J. W., Yetming, K., Patel, S. P., Auble, H., Shu, L., Ward, V. K., McCormick, C. J., Clarke, I. N., Salim, O., Wobus, C. E.,
Nawar, H. F., Lee, C. H., Connell, T. D. & other authors (2009). Thackray, L. B., Virgin, H. W., IV & Lambden, P. R. (2007). Recovery of
Ganglioside-linked terminal sialic acid moieties on murine macro- infectious murine norovirus using pol II-driven expression of full-
phages function as attachment receptors for murine noroviruses. length cDNA. Proc Natl Acad Sci U S A 104, 11050–11055.
J Virol 83, 4092–4101. Wells, S. E., Hillner, P. E., Vale, R. D. & Sachs, A. B. (1998).
Taube, S., Perry, J. W., McGreevy, E., Yetming, K., Perkins, C., Circularization of mRNA by eukaryotic translation initiation factors.
Henderson, K. & Wobus, C. E. (2012). Murine noroviruses bind Mol Cell 2, 135–140.
glycolipid and glycoprotein attachment receptors in a strain- Willcocks, M. M., Carter, M. J. & Roberts, L. O. (2004). Cleavage of
dependent manner. J Virol 86, 5584–5593. eukaryotic initiation factor eIF4G and inhibition of host-cell protein
Taube, S., Kolawole, A. O., Höhne, M., Wilkinson, J. E., Handley, S. A., synthesis during feline calicivirus infection. J Gen Virol 85, 1125–
Perry, J. W., Thackray, L. B., Akkina, R. & Wobus, C. E. (2013). A 1130.
mouse model for human norovirus. MBio 4, e00450-13. Wobus, C. E., Karst, S. M., Thackray, L. B., Chang, K.-O., Sosnovtsev,
Thackray, L. B., Duan, E., Lazear, H. M., Kambal, A., Schreiber, R. D., S. V., Belliot, G., Krug, A., Mackenzie, J. M., Green, K. Y. & Virgin,
Diamond, M. S. & Virgin, H. W. (2012). Critical role for interferon H. W. (2004). Replication of Norovirus in cell culture reveals a tropism
regulatory factor 3 (IRF-3) and IRF-7 in type I interferon-mediated for dendritic cells and macrophages. PLoS Biol 2, e432.
control of murine norovirus replication. J Virol 86, 13515–13523. Yunus, M. A., Chung, L. M. W., Chaudhry, Y., Bailey, D. & Goodfellow,
Thorne, L., Bailey, D. & Goodfellow, I. (2012). High-resolution I. (2010). Development of an optimized RNA-based murine norovirus
functional profiling of the norovirus genome. J Virol 86, 11441– reverse genetics system. J Virol Methods 169, 112–118.
11456. Zeitler, C. E., Estes, M. K. & Venkataram Prasad, B. V. (2006). X-ray
Troeger, H., Loddenkemper, C., Schneider, T., Schreier, E., Epple, crystallographic structure of the Norwalk virus protease at 1.5-Å
H.-J., Zeitz, M., Fromm, M. & Schulzke, J.-D. (2009). Structural and resolution. J Virol 80, 5050–5058.

Downloaded from www.microbiologyresearch.org by


http://vir.sgmjournals.org 291
IP: 192.188.59.82
On: Tue, 04 Jun 2019 15:38:28

You might also like