Sei sulla pagina 1di 15

Virus Research 101 (2004) 67–81

Emerging themes in rotavirus cell entry, genome organization,


transcription and replication
Hariharan Jayaram a , M.K. Estes b , B.V. Venkataram Prasad a,c,∗
a Program in Structural and Computational Biology and Molecular Biophysics, Houston, TX 77030, USA
b Department of Molecular Virology and Microbiology, Houston, TX 77030, USA
c Verna and Marrs McLean Department of Biochemistry and Molecular Biology,

Baylor College of Medicine, Houston, TX 77030, USA

Abstract

Rotaviruses, causative agents of gastroenteritis in young animals and humans, are large icosahedral viruses with a complex architecture. The
double-stranded RNA (dsRNA) genome composed of 11 segments, which codes for 6 structural and 6 non-structural proteins, is enclosed within
three concentric capsid layers. In addition to facilitating host-specific interactions, the design of the capsid architecture in rotaviruses as in
other dsRNA viruses should also be conducive to the requirement of transcribing the enclosed genome segments repeatedly and simultaneously
within the capsid interior. Several non-structural proteins facilitate the subsequent processes of genome replication and packaging. Electron
cryomicroscopy studies of intact virions, recombinant virus-like particles, functional complexes, together with recent X-ray crystallographic
studies on rotavirus proteins have provided structural insights into the capsid architecture, genome organization, antibody interaction, cell
entry, trypsin-enhanced infectivity, endogenous transcription and replication. These studies underscore contrasting features and unifying
themes between rotavirus and other dsRNA viruses.
© 2003 Published by Elsevier B.V.

Keywords: Rotavirus; dsRNA virus; Cryo-EM; Genome organization; Transcription; Replication

1. Introduction layers yield the complete virus particles (Estes, 2001).


From such studies on rotavirus and other dsRNA viruses
Rotavirus, a member of the family Reoviridae is a non- including other members of the family Reoviridae, com-
enveloped, icosahedral, double-stranded RNA (dsRNA) mon themes are emerging regarding the role of the viral
virus (reviewed in Lawton et al., 2000). The rotavirus capsid and non-structural proteins during the virus life
capsid, like all virus capsids is structured to protect its cycle.
genome and deliver it successfully into a suitable host cell, The dsRNA viruses, classified into five major groups,
in which the genome is replicated and the virus particle range from the relatively simple viruses with a single
makes copies of itself. Several recent structural and bio- dsRNA segment like in the members of the Totiviridae fam-
chemical studies have provided crucial insights into how ily to more complex viruses in the Reoviridae family, which
the different proteins encoded by the virus enable it to contain 10–12 dsRNA segments (Lawton et al., 2000; van
(1) transcribe its genome from within the confines of the Regenmortel et al., 2000). Other families of dsRNA viruses
capsid, (2) control the translation of the host genes to en- like Birnaviridae and Cystoviridae contain two and three
hance translation of the rotaviral genes in the cytoplasm dsRNA segments, respectively. Accordingly, there is a con-
of the infected host cell, (3) enable genome replication siderable variation in the complexity of the capsid organi-
wherein the negative strand is synthesized using the tran- zation in dsRNA viruses. In the members of Totiviridae and
scribed mRNA as a template, and finally (4) package Birnaviridae, a single-layer capsid encloses the genome,
the replicated dsRNA genome and assemble the capsid whereas viruses in Cystoviridae contain a double-layered
capsid. In some genera of Reoviridae such as Rotavirus,
∗ Corresponding author. Tel.: +1-713-798-5686; Orbivirus, and Reovirus, viruses exhibit distinct triple-
fax: +1-713-798-1625. layered capsid architecture. In all the well-characterized
E-mail address: vprasad@bcm.tmc.edu (B.V.V. Prasad). dsRNA viruses, the capsid structure is based on icosahedral

0168-1702/$ – see front matter © 2003 Published by Elsevier B.V.


doi:10.1016/j.virusres.2003.12.007
68 H. Jayaram et al. / Virus Research 101 (2004) 67–81

symmetry. With the exception of cystoviruses (Mindich, 2. Rotavirus capsid structure and function
1988), which use bacteria as hosts, the dsRNA viruses are
non-enveloped. Rotavirus is the major causative agent of infantile diar-
The dsRNA genome in rotavirus and other dsRNA viruses rhea accounting for nearly one million deaths annually in
presents a unique set of conditions for their survival. Be- the world (Cohen, 2001). This fact has led to the virus be-
cause the host cells do not possess the enzymatic machinery ing extensively studied to understand its pathogenicity and
to convert dsRNA into a translatable mRNA molecule, discover ways to treat and manage rotavirus infections. The
these viruses have to provide for a mechanism to synthesize rotavirus is a relatively large, ∼1000 Å in diameter, icosa-
mRNA from their genomic dsRNA. Both the transcription hedral virus. Its capsid encloses 11 segments of dsRNA
and the subsequent synthesis of progeny dsRNA have to (Fig. 1A), each segment codes for one protein with the ex-
be carried out in confined environments not only to avoid ception of segment 11, which codes for 2 proteins (reviewed
degradation of the genome by cellular nucleases but also to in (Estes, 2001). Of these 12 proteins, 6 are structural (VPs)
prevent unfavorable antiviral responses in the host cells that and 6 are non-structural (NSPs). The rotavirus capsid is
could be triggered by increased concentrations of dsRNA. composed of three concentric protein layers that enclose
All the dsRNA viruses encode their own enzymes necessary the genome (Fig. 1B and C; reviewed in Prasad and Estes,
for transcription and are capable of endogenously transcrib- 2000). The complete virions are called triple-layered parti-
ing their genomes. Together with this common attribute of cles (TLPs, Fig. 1B), particles that lack the outer layer are
endogenous transcription, the capsid architecture in these called the double-layered particles (DLPs, Fig. 1E) and are
viruses also has to account for host-specific interactions. non-infectious, and particles that lack the outer two layers
The dsRNA viruses are ubiquitous in nature and infect hosts are called the single-layered particles (SLPs, Fig. 1D).
that range from bacteria and fungi to species throughout
the plant and animal kingdoms (van Regenmortel et al., 2.1. The outer capsid layer and cell entry
2000). An interesting question is: how does capsid archi-
tecture in these viruses efficiently integrate the requirement The viruses in Reoviridae infect a wide variety of hosts.
of host-specific cell entry with the common requirement of Rotaviruses infect the cells of the intestinal epithelium, while
endogenous transcription? some viruses of the genus Orthoreovirus spread to the cen-
Structures of several dsRNA viruses including L-A virus tral nervous system and viruses among the Orbiviruses in-
of yeast origin from Tottiviridae family (Caston et al., 1997), fect both insect and mammalian hosts. The outermost layers
infectious bursal disease virus from Birnaviridae (Bottcher in Reoviridae are implicated in cell attachment, membrane
et al., 1997), ␾6 from Cystoviridae (Butcher et al., 1997, penetration and cell entry. In keeping with the wide host
2001), and several members of the Reoviridae representing range, the outer capsid layers among the Reoviridae show
various genera including rotavirus (Prasad and Estes, 2000; a remarkable range of diversity in terms of composition of
Tihova et al., 2001), bluetongue virus (BTV) in the Orbivirus proteins, their organization, their associated activities, and
(Grimes et al., 1997; Prasad et al., 1992), orthoreovirus the mechanisms of cell entry.
(Dryden et al., 1993, 1998), aquareovirus (Nason et al., 2000; Cryo-EM reconstructions of rotavirus (Prasad et al.,
Shaw et al., 1996), rice dwarf virus (RDV) in the Phytore- 1988; Prasad and Estes, 2000; Tihova et al., 2001; Yeager
ovirus genera (Zhou et al., 2001), and cypovirus (Hill et al., et al., 1990), orthoreovirus (Dryden et al., 1993, 1998),
1999; Xia et al., 2003), have been analyzed by electron cry- BTV (Grimes et al., 1997; Hewat et al., 1992; Schoehn
omicroscopy (cryo-EM) techniques. X-ray structures of L-A et al., 1997), rice dwarf virus (Lu et al., 1998) and cy-
virus (Naitow et al., 2002), and transcriptionally competent povirus (Hill et al., 1999; Xia et al., 2003), have provided
cores of BTV (Grimes et al., 1998), orthoreovirus (Reinisch structural insights into the organization of this outermost
et al., 2000), and RDV (Naitow et al., 2002), have been de- capsid layer. Accordingly, members of the Reoviridae can
termined to near 3 Å resolution. In addition to the structures be classified into two types, those that have a complete
of viral capsids, in some cases, individual viral proteins have T = 13 outer layer or the non-turreted viruses and those
been determined by X-ray crystallography (Butcher et al., that have an incomplete T = 13 layer where the normal
2001; Deo et al., 2002; Dormitzer et al., 2002; Jayaram et al., T = 13 arrangement is interrupted by the presence of a
2002; Liemann et al., 2002; Mathieu et al., 2001; Tao et al., turret-like structure at the five-fold axes of symmetry.
2002). Together these structural studies have begun to pro- In rotavirus, the outermost layer is made up of 780 copies
vide a detailed description of the capsid organization and of VP7 (38 kDa) a glycoprotein, and 120 copies of the
molecular insights into the mechanisms of various functional spike protein VP4 (88 kDa) (Prasad et al., 1990) (Fig. 1B).
activities of these dsRNA viruses. This review primarily fo- Cryo-EM studies indicate that the 780 copies of VP7 are ar-
cuses on the recent structural findings in rotavirus, and in ranged as 260 trimers that are located at the local and strict
relation to other dsRNA viruses, attempts to summarize this three-fold axes of a T = 13 (left-handed) icosahedral lattice.
information to underscore unifying principles in capsid ar- A distinctive feature of the rotavirus structure is the pres-
chitecture, genome organization, endogenous transcription ence of 132 aqueous channels, ∼140 Å deep, spanning the
and replication. outer two capsid layers, at all the five- and six-coordinated
H. Jayaram et al. / Virus Research 101 (2004) 67–81 69

Fig. 1. Architectural features of rotavirus. (A) PAGE gel showing, 11 dsRNA segments comprising the rotavirus genome. The gene segments are numbered
on the left and the proteins they encode are indicated on the right. (B) Cryo-EM reconstruction of the rotavirus triple-layered particle. The spike proteins
VP4 is colored in orange and the outermost VP7 layer in yellow. (C) A cutaway view of the rotavirus TLP showing the inner VP6 (blue) and VP2 (green)
layers and the transcriptional enzymes (shown in red) anchored to the VP2 layer at the five-fold axes. (D) Schematic depiction of genome organization
in rotavirus. The genome segments are represented as inverted conical spirals surrounding the transcription enzymes (shown as red balls) inside the VP2
layer in green. (E and F) Model from Cryo-EM reconstruction of transcribing DLPs. The endogenous transcription results in the simultaneous release of
the transcribed mRNA from channels located at the five-fold vertex of the icosahedral DLP.

positions of the T = 13 lattice. Inserted into the channels more readily and rapidly than those particles that are not
(type II channels) that surround the 12 icosahedral five-fold trypsinized (Kaljot et al., 1988; Keljo et al., 1988). In vitro
vertices are the 60 bilobed spikes, each 120 Å long. The experiments have shown that proteolytically activated parti-
molecular identity of these spikes as dimers of VP4 was cles, as well as recombinant VP5∗ , possess lipophilic activity
shown by cryo-EM reconstructions of rotavirus complexed (Dowling et al., 2000; Nandi et al., 1992; Ruiz et al., 1994).
with anti-VP4 antibodies (Prasad et al., 1990; Tihova et al., It is interesting to note here that VP4 contains a putative
2001). Subsequent cryo-EM studies discovered that VP4 has fusion domain similar to that seen in the enveloped viruses
a large globular domain that is buried inside the inner layer, such as alphaviruses and influenza viruses. Proteolysis thus
making the total length of the spike 200 Å (Shaw et al., 1993; constitutes a key process in the efficient internalization of
Yeager et al., 1994). rotaviruses into cells. This is particularly relevant consid-
ering that rotavirus replication takes place in enterocytes in
2.1.1. VP4 the small intestine, an environment rich in proteases.
Although earlier studies implicated VP7 in the cell entry The molecular mechanisms of increased infectivity by
process (Fukuhara et al., 1988; Sabara et al., 1985), sub- proteolysis and its role in the membrane penetration are not
sequent studies have increasingly indicated that VP4 is the well understood. Recent studies comparing the biochemical
major player in this process. VP4 is implicated not only in and structural properties of rotavirus grown in the presence
cell attachment and cell penetration but also in hemagglu- and absence of trypsin have provided some exciting and
tination, neutralization and virulence (Estes, 2001). VP4 is novel insights into this process, and indicated that trypsin
susceptible to proteolysis. Proteolytic cleavage of VP4 en- cleavage stabilizes the spike assembly and confers icosahe-
hances viral infectivity by several fold (Arias et al., 1996; dral ordering (Crawford et al., 2001). These results are con-
Estes et al., 1981) and facilitates virus entry into cells sistent with biochemical studies on recombinant VP4, which
(Kaljot et al., 1988). During proteolysis, VP4 (88 kDa) is show that proteolysis of monomeric VP4 yields dimeric
cleaved into VP8∗ (28 kDa, aa 1–247) and VP5∗ (60 kDa, VP5∗ (Dormitzer et al., 2001). Based on these studies it ap-
aa 248–776) and the cleavage products remain associated in pears that that trypsin cleavage has an intracellular role in
the virion (Fiore et al., 1991). Trypsinized viruses enter cells ensuring the correct conformation and proper assembly of
70 H. Jayaram et al. / Virus Research 101 (2004) 67–81

the spikes. Further structural and biochemical studies are the multi-step entry process, the conformational adaptability
required to provide a better understanding of how trypsin and flexibility of VP4 may play an important role. Studies
affects intracellular spike assembly. on rotaviruses grown in the absence and presence of trypsin
Several recent studies have shown that rotavirus cell entry together with the observation that VP4 can undergo large
is a multi-step process involving sialic acid (SA)-containing pH-induced conformational changes are consistent with such
receptors in the initial cell attachment step and integrins such a notion (Crawford et al., 2001; Pesavento et al., 2001). A
as ␣v␤3, ␣4␤1, ␣2␤1 during the subsequent post-attachment detailed picture will emerge once more intermediates in the
steps (Coulson et al., 1997; Guerrero et al., 2000; Hewish rotavirus entry process are characterized structurally.
et al., 2000; Zarate et al., 2000). In this process, the VP8∗
domain is involved in the interactions with SA, whereas 2.1.2. VP7
VP5∗ is implicated in interactions with integrins. Involve- The precise role of VP7 during early interactions of the
ment of SA during rotavirus infections is not an essential step virus with the cell is not clear, but its has been postulated that
for all rotavirus strains. In the majority of rotavirus strains, VP7 may modulate the function of VP4 during the attach-
including human rotaviruses, cell entry is SA independent ment and entry process (Beisner et al., 1998; Mendez et al.,
(Ciarlet et al., 2001). In these viruses, the majority of neu- 1996), and may interact with cell surface molecules after the
tralizing monoclonal antibodies that recognize VP4 select interaction is initiated by VP4 (Mendez et al., 1999). One of
mutations in VP5∗ , suggesting that cell entry is mediated the first events upon cell entry is the loss of this outer layer
mainly by the VP5∗ (Kirkwood et al., 1996; Kobayashi et al., to expose the transcriptionally active DLP to the cytoplasm.
1990; Padilla-Noriega et al., 1995). Studies with polarized Biochemical studies indicate that the in vivo decapsidation
epithelial cell lines show that with SA-dependent viruses, can be mimicked by treating TLPs with calcium chelators
the viral entry is restricted to the apical membrane, whereas like EDTA (Cohen et al., 1979). VP7 binds calcium and
SA-independent viruses enter either apically or basolater- the sensitivity of virions to low calcium concentrations is
ally thus further documenting variations in the entry mech- strain-dependent (Cohen et al., 1979; Gajardo et al., 1997;
anisms between SA-independent and SA-dependent viruses Ruiz et al., 1996). Several studies also have suggested
(Ciarlet et al., 2001). calcium-driven conformational changes in VP7 (Dormitzer
Recently, X-ray structure of a portion (aa 64–242) of and Greenberg, 1992). Studies on baculovirus-expressed
the sialic acid binding domain of VP8∗ was determined recombinant VP7 have shown a requirement for calcium
and this structure strongly suggests that the distal globular in the formation of VP7 trimers, which crystallize into
heads of the spikes are made of VP8∗ (Dormitzer et al., hexagonal plates mimicking the arrangement of VP7 on
2002). Interestingly, the sialic acid binding domain exhibits the capsid (Dormitzer et al., 2000). Thus, while appropriate
a ␤-sandwich fold similar to that seen in galectins, a family levels of calcium help maintain the structural integrity of
of sugar binding proteins, despite a lack of sequence similar- the VP7 layer, low calcium concentrations, similar to those
ity. Analysis of the crystal structure with bound sialic acid in the cytoplasm, trigger the disassociation of VP7 trimers
revealed that although the galectin binding site is blocked leading to uncoating of the VP7 layer. Uncoating of the
in VP8∗ , the protein binds sialic acid in a shallow groove outer layer resulting in transcriptionally competent DLPs is
on its surface using residues that are conserved in sialic a necessary event in the replication cycle of rotavirus. Some
acid-dependent rotavirus strains. A comparison of this struc- antibodies directed against VP7 neutralize the virus by in-
ture with that of the apo-form of VP8∗ , determined by NMR, hibiting the decapsidation of the TLP (Ludert et al., 2002).
suggests that the ligand binds by inducing a slight conforma- The effect of these neutralizing antibodies is not overcome
tional change in the residues involved in binding. The VP8∗ by lipofecting virus–antibody complexes into cells. In con-
structure further represents one of the first observed cases trast, some antibodies directed against VP4 neutralize by
of a protein in rotavirus and the dsRNA viruses (other than inhibiting the binding of virus to the cell, an effect that
the polymerase) taking on a fold seen among host proteins can be overcome by lipofecting these complexes into cells.
and unknown thus far among viral proteins whose struc- It is possible that neutralizing anti-VP7 antibodies prevent
tures have been determined. Based on this structural result, VP7 from undergoing necessary conformational changes
it is proposed that VP4 arose from the insertion of a host that facilitate uncoating of the VP7 layer. Thus, VP7 is the
carbohydrate-binding domain into a viral membrane inter- key mediator of the calcium driven uncoating in rotavirus
action protein (Dormitzer et al., 2002). and the function of the outer layer composed of VP4 and
The proposed sequential interactions with various lig- VP7 is to ensure that the transcriptionally active DLPs are
ands in the multi-step rotavirus entry processes may in- delivered to the host cytoplasm.
volve a series of conformational changes. For instance, in
SA-dependent strains of rotavirus, binding of sialic acid may 2.2. The intermediate layer and endogenous genome
induce conformational changes in VP8∗ to enable VP4 to transcription
bind to the integrins more efficiently. In the SA-independent
strains VP4 may already be in a conformation that is suit- The DLP resulting from the removal of the outer layer pro-
able for directly interacting with the upstream receptors. In teins, VP4 and VP7, is the transcriptionally competent form
H. Jayaram et al. / Virus Research 101 (2004) 67–81 71

of rotavirus capable of transcribing the enclosed dsRNA seg- equivalent of VP6 in rotavirus, the VP6 trimer exhibits
ments within its confines (Fig. 1E). The DLP is composed extensive lateral interactions involving charged residues.
of the remaining four structural proteins encoded by the ro- Thus, modeling has revealed that all the interacting sur-
tavirus genome of which VP6 is the major constituent. In faces (VP6–VP4, VP6–VP7, and VP6–VP2) contain the
the DLP, VP6 forms the outer most capsid layer with 260 most conserved residues of VP6. Although the base of the
trimers of VP6 assembled on a T = 13 icosahedral lattice. VP6 layer displays an overall negative surface electrostatic
In the context of the rotavirus TLP, the VP6 layer, hence- potential, the VP6 interactions with VP2 are predominantly
forth referred to as the intermediate layer, is sandwiched hydrophobic. In contrast the interactions with VP7 and VP4
between the outer T = 13 layer formed by VP7 and VP4, appear to involve several charged residues. In solution, VP6
and the inner T = 1 capsid layer formed by VP2. In the ro- forms highly stable trimers. The VP6 trimer has a bound
tavirus structure, the VP6 trimers lie below the VP7 trimers zinc ion that is shown to be necessary for the stability of
so that the aqueous channels in the two T = 13 layers are the trimer. VP6 trimers easily form 2-D crystalline arrays
in register. In the overall organization of the rotavirus, VP6 and helical tubes with pseudo-hexagonal packing, but rarely
appears to integrate the two principal functions of the virus, icosahedral shells that are similar to the T = 13 VP6 layer in
cell entry and endogenous transcription, through its interac- the rotavirus structure. The lateral interactions between VP6
tions with the outer layer proteins VP7 and VP4, and the trimers do not have all the information to form the closed
inner layer protein VP2. shell. In contrast, VP2 has the ability to form native-like
icosahedral shells of appropriate size. Therefore, it is likely
2.2.1. Pseudo-atomic model of the VP6 layer that the VP2 layer provides a proper scaffold for the assem-
Although an X-ray structure of the DLP is yet to be de- bly of VP6 trimers into a T = 13 icosahedral organization.
termined, docking of the X-ray structure of VP6 into the
cryo-EM density of the DLP has resulted in a pseudo-atomic 2.2.2. Role of VP6 in endogenous transcription
model of the T = 13 VP6 layer (Mathieu et al., 2001). VP6 Earlier biochemical studies clearly indicated that none of
has two domains, the distal domain with eight-stranded the components of the DLP alone is capable of transcribing
anti-parallel ␤-sandwich fold makes contact with the VP7 the dsRNA and that VP6, despite lack of any enzymatic
layer, and the lower domain consisting of a cluster of functions, is essential for endogenous transcription of the
␣-helices makes contact with the inner VP2 layer (Fig. 2). genome. Based on cryo-EM studies of DLPs, Prasad et al.
VP6 trimers interact laterally to form the T = 13 layer and (1988) were the first to propose that channels in the VP6
there appear to be at least two types of contacts between the layer could be used for mRNA exit. More recent cryo-EM
trimers. The contacts, across the quasi two-fold axes, closer studies on the actively transcribing DLPs, have shown that
to the icosahedral three-fold axes, are similar, whereas as of the three types of the channels in the T = 13 VP6 layer,
the trimers approach the icosahedral five-fold axis, the con- the nascent mRNA transcripts exit specifically through the
tacts are varied. In contrast to VP7 of BTV, a structural type I channels located at the five-fold axes (Fig. 1E and F)

Fig. 2. Structure of VP6 trimer. Structure of VP6 trimers showing the upper domain with its seven-strand jelly role topology and the predominantly
helical lower domain (Mathieu et al., 2001). The structure of the upper domain is analogous to the domain in BTV VP7 (Grimes et al., 1995, 1998).
The docking of the VP6 trimers into electron-density from cryo-EM reconstructions has allowed the construction of a pseudo-atomic model for the outer
surface of the DLP in rotavirus.
72 H. Jayaram et al. / Virus Research 101 (2004) 67–81

(Lawton et al., 2000). These studies also confirmed that transcription, revealed that these particles are able to initi-
DLPs maintain their structural integrity during the process of ate transcription and synthesize capped transcripts of five
transcription. In the pseudo-atomic model of the VP6 layer, to seven nucleotides in length, thereby indicating that the
a ␤-hairpin motif of VP6 with a highly conserved sequence enzymatic activities of the DLP are not affected by binding
that protrudes into the mRNA exit channel may play a of VP7 or antibody. The short length of transcripts indicates
functional role in the translocation of the nascent transcripts that elongation is not inhibited by any steric narrowing or
during endogenous transcription. The net concentration of restriction caused by antibody-binding, or analogously VP7
negative charges on the walls lining the type I channel may binding, but instead can be attributed to the subtle conforma-
facilitate the extrusion of the mRNA transcript by increasing tional changes near the VP6–VP2 interface observed consis-
its fluidity because of the electrostatic repulsion between tently in the structures of transcriptionally inactive TLP and
the interior surface of the channel and the mRNA transcript. DLP–antibody complexes. A similar cryo-EM study using
A detailed mutational analysis based on the pseudo-atomic a different set of antibodies with contrasting effects on tran-
model of the VP6 layer has helped to elucidate the deter- scription however concluded that the inhibitory effect of one
minants of VP6 required for assembly on VP2, and how of the two antibodies studied may be due to rigidification of
VP6 may affect endogenous transcription (Charpilienne the VP6 trimers upon antibody binding (Thouvenin et al.,
et al., 2002). Thirteen site-specific substitution mutations 2001). Although no conformational changes in VP6 with
of amino acids residues that directly contact the VP2 layer, either of the antibodies were seen, there were major differ-
as identified in the pseudoatomic model, have been stud- ences in the extent of interaction between VP6 and these
ied in terms of their ability to trimerize, to form virus-like antibodies. With the antibody that inhibited transcription,
particles when co-expressed with VP2, and to assemble five loops from two of the VP6 subunits in the trimer were
onto wild-type cores and form in vitro reconstituted DLPs involved in the antibody binding, whereas with the other
that could transcribe the enclosed genomes and produce antibody, which did not inhibit transcription, only one VP6
rotaviral mRNA. In cases where defects were observed, the monomer was involved. From these observations, Thouvenin
amino acids essential for recovery of transcription or as- et al. (2001) proposed that the inhibitory effect of one of the
sembly were identified. All the VP6 mutants formed stable antibodies is caused by preventing conformational changes
trimers and self assembled into tubular structures similar to in the VP6 layer required for transcription. Although the pro-
wild-type VP6. This is consistent with the observation from posed mechanisms of how a VP6-specific antibody inhibits
the X-ray structure of VP6 that the lower ␣-helical domain transcription by these two studies differ, both the studies
of VP6 does not contribute to either the trimeric interactions clearly indicate that the dynamics of the interaction between
or the lateral interactions in the assembly of the tubular the VP6 and VP2 layers are important for transcription.
structures. Of the 13 VP6 mutants examined, 3 were unable The exit of transcripts through the channels at the
to assemble with VP2 and 3 others partially assembled. five-fold axes appears to be an emerging common theme in
These mutants either did not rescue the transcriptase activ- dsRNA viruses. In orthoreovirus, both conventional EM and
ity or did so only marginally substantiating further that the cryo-EM studies have shown the transcripts exit through
proper assembly of VP6 trimers on VP2 is an absolute re- the turrets at the five-fold axes. In BTV cores, X-ray crys-
quirement for endogenous transcription. Four other mutants tallographic analysis of core crystals soaked with various
that generally preserved the hydrophobic character of the substrates and products of the transcription reaction has
contact region, assembled and transcribed well, emphasiz- shown that the mRNA exit channel in the BTV core is
ing the importance of hydrophobic interactions in the proper likely to be at the five-fold vertex (Diprose et al., 2001).
assembly and stability of VP6 on VP2. An interesting result In addition to providing conduits for the exit of transcripts,
was obtained with three of the mutants, which had an extra another important function of the VP7 layer (equivalent of
charge introduced at the VP6–VP2 interface. These mutants VP6 layer in rotavirus), that became evident from these
assembled well on cores, but surprisingly did not rescue studies, is that certain regions surrounding the five-fold axes
the transcriptase activity in the reconstituted DLPs. This in this layer function as substrate sinks for the transcription
result prompted the investigators to favor a hypothesis in reaction. Although the dynamic events behind endogenous
which transcript extrusion during endogenous transcription transcription are still unclear, the structural and biochem-
requires VP6 to undergo subtle conformational changes and ical studies on several dsRNA viruses have brought into
the extra charge in the mutants inhibits such changes. focus the elegance of capsid organization in these viruses
The idea of conformational changes in VP6 associated and reveal the fact that capsid architecture and genome or-
with endogenous transcription is also suggested by the ganization are coordinated to ensure multiple and repeated
cryo-EM structural studies on DLP-anti (VP6) MAb com- rounds of transcription.
plexes. Lawton et al. (1999) in their studies showed that
certain antibodies inhibit transcription as does the presence 2.3. The innermost layer
of the outermost VP7 layer, which renders the TLP tran-
scriptionally inactive. Biochemical analysis of TLP, and The core of the rotavirus consists of the remaining three
DLP–antibody complexes in which the antibody inhibited structural proteins VP1, VP2, and VP3 (Fig. 1C and D). Of
H. Jayaram et al. / Virus Research 101 (2004) 67–81 73

these three proteins, VP2 is the most abundant and forms the subsequently has been seen in aquareovirus, orthoreovirus,
innermost layer interacting with the VP6 layer on the out- BTV, RDV, cypovirus, and ␾6 virus. However, one prin-
side and the genomic RNA on the inside. The other two pro- cipal difference between rotavirus, BTV, or orthoreovirus,
teins VP1, an RNA-dependent RNA polymerase (Valenzuela and aquareovirus or cypovirus, is that in the latter group
et al., 1991), and VP3, guanylyl and methyl transferase of viruses the capping enzyme is external to the innermost
(Chen et al., 1999; Liu et al., 1992), are present in small layer forming a distinct turret-like structure at the five-fold
quantities and provide the enzymatic functions required for axis, although the polymerase is internal. The location of the
producing the capped mRNA transcripts. Of all the structural transcription enzyme complex at the five-fold axis is con-
proteins of rotavirus, VP2 is the only protein that has the abil- sistent with the nascent transcripts exiting through the chan-
ity to self-assemble into a native-like icosahedral structure, nels at the five-fold axes. In rotavirus, biochemical studies
when expressed in insect cells. This observation strongly on recombinant VLPs containing VP2 with amino-terminal
indicates that VP2 possesses the determinants required to deletions co-expressed with VP6, VP1 and VP3 indicate that
direct the proper assembly of other rotavirus proteins. removal of 25 N-terminal residues of VP2 completely pre-
vent the incorporation of VP1 and VP3. Thus, in addition to
2.3.1. Unique organization of the inner most capsid layer RNA-binding activity, the N-terminal residues of VP2 are in-
The VP2 layer consists of 120 molecules organized as 60 volved in anchoring VP1 and VP3. The N-terminal residues
dimers on a T = 1 icosahedral lattice. Such an icosahedral of the one of the two subunits in the icosahedral asymmetric
organization with 120 subunits is quite unique and is found unit of the “T = 2” that is close to the five-fold axis may be
only in dsRNA viruses. This is the most conserved feature involved in anchoring the transcription enzyme, whereas the
in all structurally characterized members of the Reoviridae N-terminal residues of the other subunit, which is slightly
and other dsRNA viruses such as ␾6, a bacterial virus and offset from the five-fold axis may involved in the interac-
L-A virus of yeast origin. The atomic level description of tions with the underlying genomic RNA. It is possible that
such an organization with two subunits in the icosahedral the unique organization of the innermost layer in the dsRNA
asymmetric units, and hence called a T = 2 icosahedral viruses has evolved to serve the dual purpose of properly po-
structure, is provided by the X-ray crystallographic analy- sitioning the transcription enzyme complex and organizing
ses of BTV and orthoreovirus cores (Grimes et al., 1998; the genome to facilitate endogenous transcription.
Reinisch, 2002; Reinisch et al., 2000). In these structures, Although the locations of the transcription enzymes can
one of the two subunits in the icosahedral asymmetric unit be inferred from the structures of the intact virions, the pre-
points toward the icosahedral five-fold axis and the other is cise molecular structures of these enzymes inside the virions,
slightly offset from the five-fold axis. Each subunit has three with the exception of the capping enzyme in the “turretted”
domains with the N-terminal residues facing inward toward dsRNA viruses such as orthoreovirus, are not resolved as
the virus genome. Similar subunit structure and organiza- they are present in non-icosahedral amounts. Therefore,
tion is also seen in the high-resolution cryo-EM structures structural understanding of how these enzymes function
of RDV and CPV and is likely to be replicated in rotavirus has to come from the studies on individual components.
as indicated by the cryo-EM analysis of the recombinant Recently, X-ray crystallographic structures of the ␾6 and
VP2/6 particles with full-length VP2 and a mutant VP2 with orthoreovirus polymerases have been determined (Butcher
N-terminal residues deleted. VP2 exhibits RNA-binding et al., 2001; Tao et al., 2002). These studies have provided
ability through its N-terminal residues. Consistent with this significant mechanistic insights into how these molecules
observation, several points of contact between the VP2 layer function both in the transcription and replication processes.
and RNA, presumably through the N-terminal residues of Both the structures have a canonical finger–palm–thumb
VP2, are observed in the cryo-EM structure of the DLP. core, as seen in several other polymeases, surrounded by
Similarly, interactions between VP3 and RNA are seen the N- and C-terminal elaborations. These elaborations differ
X-ray structure of the BTV core. Thus, one of the principal significantly between the two structures perhaps reflect-
functions of the innermost layer in these viruses may be ing the differences in the transcription strategies in these
to direct the structural organization of the genome that is viruses. In ␾6 and birnaviruses, which contain three and
conducive for its endogenous transcription. two dsRNA segments, respectively, the transcription is
semi-conservative. In the members of Reoviridae, which
2.3.2. Transcription enzyme complex contain 10–12 dsRNA segments, the transcription is conser-
Where are the transcription enzymes located? Using a vative. Generally, the polymerase molecules in the members
comparative cryo-EM analysis of DLP and the recombinant of Reoviridae are larger than their counterparts in other
virus-like particles with and without VP1 and VP3, Prasad dsRNA viruses. The structure of the reovirus polymerase,
et al. (1996) were the first to show that these minor proteins perhaps a better representative of the rotavirus polymerase,
were incorporated as a heterodimer anchored to the inside shows a cage-like structure with four channels leading to
surface of the VP2 layer at each of the 12 five-fold ver- the central catalytic core (Tao et al., 2002). Further X-ray
tices. A similar structural organization of the transcription crystallographic analyses with appropriate ligands have as-
enzymes anchored to the inner surface of the T = 2 layer signed these channels to specific functions such as substrate
74 H. Jayaram et al. / Virus Research 101 (2004) 67–81

and template entry, mRNA or dsRNA exit depending upon ganization of the genome in a dsRNA virus was provided
the polymerase involvement during transcription or replica- by the cryo-EM analysis of rotavirus (Prasad et al., 1996).
tion (Tao et al., 2002). Interestingly, these structural studies These studies indicated that the viral dsRNA forms a do-
also revealed a binding site for the 5 cap structure seen in decahedral structure in which the RNA double helices,
reovirus and all the members of the Reoviridae, suggesting interacting closely with the VP2 layer, are packed around
a potential mechanism for tethering of the non-template the transcription complexes located at the icosahedral ver-
strand during transcription which helps localize the tem- tices. It was suggested that VP2, which is icosahedrally
plate strand and allows for repeated cycling of the template ordered, with its RNA-binding property is responsible for
during the continuous endogenous transcription. One pri- the icosahedral ordering of the closely interacting portions
mary difference between the rotavirus and orthoreovirus is of the RNA, and this ordering is diminished at lower radii.
the location of the capping enzyme. In contrast to orthore- Subsequent X-ray crystallographic studies on BTV (Gouet
ovirus, in which the capping enzyme is located on the outer et al., 1999) and orthoreovirus cores (Reinisch et al., 2000)
surface of the innermost layer, in rotavirus, biochemical and cryo-EM studies on RDV, CPV, aquareovirus, have all
and structural studies have shown that the capping enzyme consistently shown that a significant portion of the genome
is in close proximity to the polymerase (Lawton et al., is statistically ordered and manifests as concentric layers
1999; Prasad et al., 1996). It remains to be seen how the of density in the icosahedrally averaged structures of these
polymerase in rotavirus accommodates interaction with the viruses. These concentric layers of RNA are generally
capping enzyme. separated by 28–30 Å. Assuming a local hexagonal pack-
ing of the RNA helices, such separation translates into an
inter-strand spacing of 30–32 Å. The observations in these
3. Structural organization of the genome icosahedrally-averaged structures are consistent with other
studies. A similar inter-strand spacing was deduced from
Understanding the structural organization of the genome earlier low angle X-ray scattering studies on orthoreovirus,
particularly in dsRNA viruses assumes a greater impor- which also suggested that the dsRNA genome is tightly
tance because of the intimate involvement of the genome packed as parallel helices in a semicrystalline array (Harvey
in the various enzymatic activities within the confines of et al., 1981). Considering the volume available for the
the capsid. From all the available biochemical and struc- genome inside the capsid layers and the molecular weight
tural data to date on several members of the Reoviridae, of the genome in these viruses, typically the concentration
including cypovirus, orthoreovirus, BTV and rotavirus, the of the RNA inside the capsid layer is around 400 mg/ml
emerging consensus is that an independently functioning (Gouet et al., 1999). At such concentrations, dsDNA is
transcription enzyme complex, anchored to the inside sur- known to exhibit a columnar hexagonal liquid crystalline
face of the innermost capsid layer, transcribes each genome packing with an inter-strand spacing of about ∼30–32 Å
segment, and that all the genome segments are transcribed (Livolant et al., 1989), suggesting thereby the packaged
simultaneously (Banerjee and Shatkin, 1970; Bartlett et al., dsRNA in these viruses behaves like dsDNA. How are the
1974; Gillies et al., 1971; Skehel and Joklik, 1969; Smith charges on the condensed dsRNA neutralized? Generally
and Furuichi, 1982). In vitro studies indicate that the tran- counter ions like Mg2+ , or organic polyamines like sper-
scriptionally competent particles of these viruses are highly mines or spermidines are involved in nucleic acid charge
efficient molecular machines capable of repeated cycles neutralization in viruses. However, Gouet et al. (1999),
of transcription. During each cycle of transcription, the found that the levels of metal counter ions in BTV cores,
dsRNA segment, which has to move around the anchored measured using a scanning proton microprobe, are too low
polymerase, must be unwound, separated, rejoined, and re- to be implicated in charge neutralization, and instead have
wound for further cycles of transcription. The transcription suggested that the charge neutralization may be through
process appears to be impressively fast. In orthoreoviruses, organic polyamines in this virus.
it is estimated that transcription proceeds at a rate of 50
nucleotides/s. The structural organization of the genome in 3.2. Reversible condensation and expansion of the
these viruses thus must allow for what seems to be a well rotavirus genome
orchestrated dynamic process of repeated, simultaneous,
high speed transcription of multiple segments. Recent cryo-EM analysis of rotavirus examined under
various chemical conditions unraveled a remarkable ability
3.1. Manifestation of genomic RNA in of the rotavirus genome to undergo reversible condensation
icosahedrally-averaged structures and expansion within the capsid interior (Pesavento et al.,
2001) (Fig. 3). These studies have provided further insights
Although the precise organization of the genome in the into the structural organization of the genome and the na-
members of Reoviridae remains to be elucidated, recent ture of interactions between the genome and the internal
structural studies on some of these viruses have provided proteins. At high pH in the presence of ammonium ions, the
useful insights. The first visualization of the structural or- genome condenses to a radius of 180 Å from the original
H. Jayaram et al. / Virus Research 101 (2004) 67–81 75

Fig. 3. Reversible condensation and expansion of the rotavirus genome. A cartoon representation of a plausible structural organization of the genome in
normal physiological condition (left) and in the ammonium high pH condition (right). Such a model from several biochemical and structural studies (Gouet
et al., 1999) offers a simple mechanistic explanation of the observed genome condensation in rotavirus (Pesavento et al., 2001). The condensation from a
original radius of 220–180 Å is achieved by simply reducing the inter-strand spacing in the spiral from 31 Å (indicated by pair of arrows in bottom left),
as observed in the normal physiological conditions, to a spacing of 25 Å (pair of arrows bottom right) as observed in the ammonium high pH condition.

radius of 220 Å, and when brought back to physiological pH RNA-binding property, plays an important role in maintain-
the genome expands to its original radius. The genome re- ing the appropriate spacing between the RNA strands in the
mains transcriptionally viable after returning to its original native expanded state. In the native state the genome exhibits
state, thereby indicating that this transformation is merely several points of contact with the VP2 layer. As the genome
structural and the dsRNA is not covalently disrupted by proceeds to the condensed state, these contacts are broken
the ammonium high pH treatment. Apart from VP4, which only to be reestablished when the physiological conditions
is irreversibly altered, the rest of the capsid remains intact are restored. The observed condensation is a synergetic ef-
under these conditions. These studies illustrate the remark- fect of hydroxyl and ammonium ions, as it is not observed
able stability of the capsid and resilience of the genome; at high pH alone, or with increased concentrations of NH+
these attributes may be required to carry out the continuous or Mg2+ ions at physiological pH. The two inferences that
transcription of multiple segments within the capsid. can be made from these observations are that VP2–RNA in-
Several observations made in this study have direct rel- teractions are pH-dependent and that the negative charges
evance to the roles of VP2, and the transcription enzyme on genomic RNA in the native state may only be partially
complexes in the structural organization of the genome. The neutralized. The effect of high pH is to disrupt RNA–VP2
observation that the condensation of the genome is isometric interactions, possibly due to deprotonation of critical VP2
and concentric with respect to the particle center, suggests residues, and cause conformational changes in the RNA to
that strong interactions between the genome and the symmet- allow further charge neutralization by ammonium ions to
rically disposed transcriptional complexes exist, and such cause condensation. The partial neutralization of the ge-
interactions may play a critical role in the structural organi- nomic RNA in the native state may be necessary to maintain
zation of the genome. In the condensed state, the reconstruc- appropriate inter-strand spacing to allow the genomic RNA
tions show strong density linking the RNA core with the in- to move around the transcription complex during transcrip-
side surface of the VP2 layer at all the five-fold vertices, like tion. In the condensed state because of the reduced volume
spokes linking a hub and wheel. Without such symmetrically and consequent increase in the concentration of RNA, the
disposed interactions, the condensed RNA core once disas- inter-strand spacing is reduced from 30 to 25 Å.
sociated from the VP2 layer would not have been constrained
to remain at the particle center. The isometric condensation 3.3. A model for structural organization of the genome
also supports the notion that each dsRNA segment may be
associated with a transcription complex and all the segments Any model for the structural organization of the genome
are similarly affected by the ammonium high pH treatment. particularly in the members of Reoviridae should allow for
The studies by Pesavento et al. (2001) also underscore simultaneous, independent and repeated transcription of the
the importance of VP2 in the structural organization of the genome segments. Studies by Pesavento et al. (2001) impose
genome. These studies suggest that the VP2, through its further constrains of reversible condensation and expansion
76 H. Jayaram et al. / Virus Research 101 (2004) 67–81

on such a model. From the elegant X-ray crystallographic the dsRNA viruses, each virus particle is thought to contain
analysis of the BTV cores, which provide a more detailed a full complement of the genome. The only known excep-
and discernible electron density distribution of the genomic tion is chrysovirus, in which the genome consisting of four
RNA, particularly after including lower resolution diffrac- dsRNA segments are packaged separately into four particles
tion data and reducing the noise using real space Gaussian (Wood and Bozarth, 1972). In multi-stranded dsRNA viruses
filtering, Gouet et al. (1999) proposed a plausible model such as rotavirus, BTV, and reovirus, it remains a mystery
for the arrangement of dsRNA segments in the BTV core. as to how each particle procures a correct set of dsRNA seg-
In this model, each dsRNA segment is spooled around a ments. Although the molecular mechanisms are presently
transcription enzyme complex at the five-fold vertex. This unclear, it is evident that the virus-encoded non-structural
model allows for up to 12 independent transcription com- proteins play a major role in choreographing the entire pro-
plexes, each attached to an individual dsRNA segment for cess of genome replication, packaging, and perhaps segment
concurrent transcription. Consistent with this idea, to date assortment in these viruses. Recently, structures of two of the
no dsRNA virus with more 12 segments has been observed. non-structural proteins encoded by rotavirus have been de-
In addition to allowing for the observed simultaneous, re- termined. These structural studies have begun to shed some
peated, and independent transcription of the genome, this light into the molecular mechanisms of genome replication
model also appears to be consistent with the isometric and and packaging.
concentric condensation of the genome observed in rotavirus
as shown schematically in Fig. 3. 4.1. Structural studies on NSP3

In rotavirus, translation of the viral mRNA transcripts


4. Genome replication and packaging is facilitated by NSP3, a basic 63 kDa protein that recog-
nizes the consensus 3 sequence on the viral transcripts. As
One of the least understood processes in the rotavirus mentioned above, rotaviral transcripts are capped at their
replication cycle is the genome replication and packaging. 5 end by the action of the structural protein VP3 during
This also is true with other members of the Reoviridae. transcription but their 3 ends are not poly-adenylated. Ro-
Our understanding of the molecular basis of these processes taviruses rely on the host translation machinery to produce
particularly in the members of Reoviridae has been limited the proteins encoded by the genome. In the host cells,
mainly because of lack of a reverse genetics system. In two only poly-adenylated and capped messages are efficiently
of the dsRNA viruses, ␾6 (Mindich, 1999) and birnavirus translated. This is brought about by recognition of the 5
(Mundt and Vakharia, 1996), which contain three and two cap by eIF4E and the poly-A tail by PABP (poly-A bind-
dsRNA segments, respectively, a reverse genetics system has ing protein) which then interacts with a cellular factor
been successfully established. However, several in vivo and eIF4G, a multipurpose adaptor protein that is responsible
in vitro studies on rotavirus, reovirus, and BTV have pro- for delivering capped and poly-adenylated messages to
vided significant information on the downstream events tak- the ribosome. Rotaviruses overcome the lack of a poly-A
ing place following endogenous transcription. Subsequent to tail, which would hamper their efficient translation, by us-
endogenous transcription and release of the transcripts, the ing a consensus sequence at their 3 ends that specifically
rotavirus replication cycle may be viewed as having three binds virus-encoded NSP3. While the N-terminal domain
major stages: (1) translation and synthesis of the viral pro- of NSP3 binds this consensus sequence, the C-terminal half
teins; (2) replication, genome packaging and DLP assem- interacts with eIF4G with an affinity greater than PABP
bly; (3) budding of the newly formed DLPs into the ER to give translation of the rotavirus messages a selective
and assembly of the outer layer to form mature TLPs (re- boost following infection. Thus, NSP3 is a protein en-
viewed in (Estes et al., 2001). The capped positive-stranded coded by the virus to subvert the host translation machinery
RNA transcripts encode the rotaviral proteins and function and selectively enhance the translation of virally encoded
as templates for production of negative strands to make the mRNA. How NSP3 achieves this was demonstrated by the
progeny dsRNA. The synthesis of the negative strand and the X-ray structure of the two domains of NSP3 bound to the
subsequent duplex formation is facilitated by the viral poly- consensus 3 mRNA sequence (Deo et al., 2002) and the
merase VP1. The genome transcription and the replication structure of the other NSP3 domain bound to a peptide cor-
are thus complimentary processes both involving the viral responding to the binding site on eIF4G (Groft and Burley,
polymerase. Just as transcription takes place in the confined 2002). The asymmetric NSP3 homo-dimer with an unusu-
environment of the capsid interior, the genome replication ally large dimeric interface binds the 3 consensus sequence
also appears to take place inside a protected environment (5 -GUGACC-3 ) and completely buries most of it within a
(Patton and Spencer, 2000). In none of the dsRNA viruses basic deep cleft on the surface of the homo-dimer creating
has free dsRNA been found in infected cells. a dead end for the 3 terminal nucleotides (Fig. 4). This
Genome replication and packaging in rotavirus and the tight interaction not only promotes translation of the ro-
viruses in the Reoviridae, take place within the cytoplasmic tavirus mRNA but also prevents degradation of the rotavirus
inclusions called viroplasms (Petrie et al., 1984). In most of message by cellular nucleases. The NSP3 homo-dimer
H. Jayaram et al. / Virus Research 101 (2004) 67–81 77

and are the main constituents of the replication interme-


diates (Aponte et al., 1996; Gallegos and Patton, 1989).
Biochemical studies on recombinant NSP2 have shown
that it readily forms octamers and has NTPase (nucleotide
triphosphatase), RNA-binding and nucleic acid helix desta-
bilizing activities (Taraporewala et al., 1999; Taraporewala
and Patton, 2001). Based on these studies, it is hypothesized
that NSP2 may function as a molecular motor to facilitate
genome packaging using the energy derived from NTP
hydrolysis (Taraporewala et al., 1999).
Recently, the X-ray structure of NSP2 to a resolution of
2.6 Å has been determined (Fig. 5). These studies have pro-
vided a firm ground to begin developing a mechanistic un-
derstanding of how NSP2, in concert with NSP5 and VP1,
may facilitate genome replication and packaging. NSP2
crystallizes as an octamer using the crystallographic 4–2–2
Fig. 4. NSP3 structure. Structure of the rotavirus NSP3 N-terminal domain symmetry, with one monomer per asymmetric unit. The
in complex with a 3 consensus sequence of rotaviral mRNA (Deo et al., intrinsic ability of NSP2 to form octamers in solution is sup-
2002). A single 5 -GUGACC-3 RNA segment is buried within a basic ported by other biophysical studies including cryo-EM stud-
deep tunnel formed in the asymmetric dimer. Each subunit of the dimer
participates in different interactions with the mRNA segment.
ies. In vivo studies also strongly suggest that the oligomeric
structure is the functional form of the NSP2 (Kattoura et al.,
1994).
is thought to co-fold with the 3 consensus end into this NSP2 is a two-domain protein, which can be classi-
extremely stable complex, which increases the stability of fied as an ␣/␤ protein based on the observed secondary
the NSP3 hetero-dimer. These studies further underscore the structures. A characteristic feature in the NSP2 monomer
functional significance of the conserved sequences at the ter- is a 25 Å deep cleft between the N- and C-terminal do-
mini of the transcripts in genome translation via their recog- mains of the protein. The N-terminal domain, predomi-
nition by NSP3 (Poncet et al., 1994; Wentz et al., 1996). nantly made of ␣-helices, exhibits a novel fold, whereas
the C-terminal domain, surprisingly, despite any notice-
4.2. Structural studies on NSP2 able sequence homology, exhibits a fold that is observed
in the Histidine Triad (HIT) proteins, a family of ubiqui-
Several in vivo and in vitro studies on rotavirus have tous cellular proteins that hydrolyze nucleotides (Brenner,
strongly implicated two of the non-structural proteins NSP2 2002; Lima et al., 1997; Lima et al., 1996). This struc-
and NSP5 in genome replication and packaging. In vivo tural similarity with the HIT proteins led to the proposal
studies have shown that these two proteins along with VP1, that the cleft might correspond to the active site for NTP
the RNA polymerase, are co-localized in the viroplasms hydrolysis.

Fig. 5. NSP2 Structure. Structure of the NSP2 monomer (left), showing a deep cleft (arrow) which may be the site for NTP binding (Jayaram et al., 2002).
Structure of the functional NSP2 octamer showing a view down the four-fold (middle) and a view down one of the two-fold axis in octamer (right). The
deep grooves shown (right, arrows) are lined by basic residues and may be the sites for binding viral mRNA during genome replication and packaging.
78 H. Jayaram et al. / Virus Research 101 (2004) 67–81

The nucleic acid binding activity of the NSP2 is inde- by non-structural proteins including co-assembly of core
pendent of the NTPase activity unlike in a classical helicase proteins and genome segments remain a distinct possibility.
where the two activities are coupled. In that respect, NSP2
cannot be described as a helicase. In many of the oligomeric
helicases, nucleic acid binds inside a large central hole (Yu 5. Concluding remarks
et al., 1996). The donut-shaped NSP2 octamer also displays
a central hole with a diameter large enough for binding Despite the lack of a reverse genetics system, in recent
nucleic acid. However, both the rim and interior surface of years significant progress has been made in our understand-
the hole are mostly hydrophobic, and thus it is unlikely that ing of the structure–function relationships in rotavirus. This
the nucleic acid binds inside the hole in the NSP2 octamer. has been primarily due to the advances in molecular bi-
Instead, the prominent grooves, lined by positively-charged ology of rotaviruses that have resulted in cloned rotaviral
residues at the sides of the octamer are possible locations genes and purification of protein complexes and virus-like
for nucleic acid binding. The proposed NTP binding sites particles. Starting with the first cryo-EM reconstruction of
in the monomeric subunits are located on either side of rotavirus in 1988, which provided a low resolution picture
this groove. Although NTP hydrolysis is not directly linked of the virus architecture (Prasad et al., 1988) subsequent
either to the RNA binding or to the helix destabilizing ac- cryo-EM studies on virus–antibody complexes, recombi-
tivity of NSP2, it is possible that binding of nucleotide into nant virus particles (Prasad et al., 1990, 1996), and more
the cleft alters the conformation of the monomer and affects recent X-ray crystallographic analyses of several rotaviral
octamer–RNA interactions. Such allosteric changes may proteins have enabled a better understanding of the molec-
facilitate translocation of the bound RNA through the poly- ular mechanisms underlying cell entry (Dormitzer et al.,
merase during replication resulting in concomitant synthesis 2002), antibody neutralization (Prasad et al., 1990; Tihova
and packaging of duplex RNA into assembling particles. et al., 2001), trypsin-enhanced infectivity (Crawford et al.,
Indeed the octameric structure of NSP2 makes it a tempt- 2001), assembly (Mathieu et al., 2001), genome organiza-
ing platform or a scaffold around which the replication tion (Pesavento et al., 2001; Prasad et al., 1996), endogenous
complex is organized. It is possible that the hydrophobic transcription (Lawton et al., 1997, 1999), mRNA translation
side of the octamer, around the four-fold axis, may bind (Deo et al., 2002), and genome replication and packaging
to the VP1; given that NSP5 is an acidic protein, the basic (Jayaram et al., 2002). The outcome of such studies should
grooves of the NSP2 octamer may be the sites for binding be potentially useful in the development of vaccines and
NSP5. Although the role of NSP5 in the overall process of identification of suitable targets for rational drug design to
replication remains to be elucidated, it is plausible, that by counteract rotavirus.
having its binding site on NSP2 overlap with that of its RNA In parallel, cryo-EM and X-ray crystallographic studies
binding site, the function of NSP5 is to regulate the binding on other dsRNA viruses have underscored how the mod-
of nucleic acid by NSP2 during replication and packaging. ular architecture of the virus can integrate disparate host
In other members of Reoviridae, the existing biochem- specificities with the common requirement of transcribing
ical data suggest that NSP2 may be functionally homol- the dsRNA segments within the capsid interior. Despite the
ogous to NS2 of BTV and sNS of reovirus (Fillmore lack of extensive sequence identity, the structural and bio-
et al., 2002; Gillian and Nibert, 1998; Gillian et al., 2000; chemical studies thus far indicate a remarkable convergence
Taraporewala et al., 2001). Another protein with which ro- in both form and function necessitated by the endogenous
tavirus NSP2 may have similarities is P4 of ␾6, which is an transcription. Further biochemical and structural studies are
NTPase (Gottlieb et al., 1992). However, one critical differ- required to establish whether these viruses share any com-
ence is that, unlike rotavirus NSP2 and its putative counter- mon themes in the events downstream of transcription. A
parts in BTV and reovirus, P4 is a structural protein. It is an major achievement in the near future may be the establish-
integral part of the transcriptionally competent ␾6 core struc- ment of reverse genetics systems for some of the members
ture, located at each of the five-fold vertex as a hexamer (de of Reoviridae. Recent developments in the use of RNA in-
Haas et al., 1999). Because of the well-established in vitro terference techniques are exciting and likely to provide fur-
replication and packaging system, ␾6 is the only dsRNA ther insights into the specific role of each of the virally
virus for which there is more definitive understanding of encoded protein in the pathogenesis and morphogenesis of
how the genomic dsRNA segments are packaged inside these viruses (Dector et al., 2002).
(Mindich, 1999). Assisted by the packaging protein P4, the
three dsRNA segments in ␾6 are packaged sequentially into
a preformed core, which undergoes significant conforma- Acknowledgements
tional change upon packaging. It remains to be seen whether
such a model for packaging RNA into preformed cores is ap- This work is supported by grants from the Robert Welch
plicable to dsRNA viruses with larger numbers of segments. Foundation (BVVP), National Institutes of Health AI 36040
Based on the existing biochemical and structural data on (BVVP) and DK 31044 (MKE). We thank Drs. J.A. Lawton
Reoviridae members, alternative packaging models assisted and J.B. Pesavento for help with the figures.
H. Jayaram et al. / Virus Research 101 (2004) 67–81 79

References viral transcription complex: the bluetongue virus core. EMBO J. 20,
7229–7239.
Dormitzer, P.R., Greenberg, H.B., 1992. Calcium chelation induces a con-
Aponte, C., Poncet, D., Cohen, J., 1996. Recovery and characterization of
formational change in recombinant herpes simplex virus-1-expressed
a replicase complex in rotavirus-infected cells by using a monoclonal
rotavirus VP7. Virology 189, 828–832.
antibody against NSP2. J. Virol. 70, 985–991.
Dormitzer, P.R., Greenberg, H.B., Harrison, S.C., 2000. Purified recom-
Arias, C.F., Romero, P., Alvarez, V., Lopez, S., 1996. Trypsin activation
binant rotavirus VP7 forms soluble, calcium-dependent trimers. Vi-
pathway of rotavirus infectivity. J. Virol. 70, 5832–5839.
rology 277, 420–428.
Banerjee, A.K., Shatkin, A.J., 1970. Transcription in vitro by
Dormitzer, P.R., Greenberg, H.B., Harrison, S.C., 2001. Proteolysis of
reovirus-associated ribonucleic acid-dependent polymerase. J. Virol.
monomeric recombinant rotavirus VP4 yields an oligomeric VP5, core.
6, 1–11.
J. Virol. 75, 7339–7350.
Bartlett, N.M., Gillies, S.C., Bullivant, S., Bellamy, A.R., 1974. Electron Dormitzer, P.R., Sun, Z.Y., Wagner, G., Harrison, S.C., 2002. The rhesus
microscopy study of reovirus reaction cores. J. Virol. 14, 315–326. rotavirus VP4 sialic acid binding domain has a galectin fold with a
Beisner, B., Kool, D., Marich, A., Holmes, I.H., 1998. Characterisation of novel carbohydrate binding site. EMBO J. 21, 885–897.
G serotype dependent non-antibody inhibitors of rotavirus in normal Dowling, W., Denisova, E., LaMonica, R., Mackow, E.R., 2000. Selective
mouse serum. Arch. Virol. 143, 1277–1294. membrane permeabilization by the rotavirus VP5, protein is abrogated
Bottcher, B., Kiselev, N.A., Stel’Mashchuk, V.Y., Perevozc hikova, N.A., by mutations in an internal hydrophobic domain. J. Virol. 74, 6368–
Borisov, A.V., Crowther, R.A., 1997. Three-dimensional structure of 6376.
infectious bursal disease virus determined by electron cryomicroscopy. Dryden, K.A., Farsetta, D.L., Wang, G., Keegan, J.M., Fields, B.N., Baker,
J. Virol. 71, 325–330. T.S., Nibert, M.L., 1998. Internal/structures containing transcriptase-
Brenner, C., 2002. Hint, Fhit, and GalT: function, structure, evolution, related proteins in top component particles of mammalian orthore-
and mechanism of three branches of the histidine triad superfamily ovirus. Virology 245, 33–46.
of nucleotide hydrolases and transferases. Biochemistry 41, 9003– Dryden, K.A., Wang, G., Yeager, M., Nibert, M.L., Coombs, K.M.,
9014. Furlong, D.B., Fields, B.N., Baker, T.S., 1993. Early steps in re-
Butcher, S.J., Dokland, T., Ojala, P.M., Bamford, D.H., Fuller, S.D., 1997. ovirus infection are associated with dramatic changes in supramolec-
Intermediates in the assembly pathway of the double-stranded RNA ular structure and protein conformation: analysis of virions and
virus phi6. EMBO J. 16, 4477–4487. subviral particles by cryoelectron microscopy and image reconstruc-
Butcher, S.J., Grimes, J.M., Makeyev, E.V., Bamford, D.H., Stuart, D.I., tion. J. Cell Biol. 122, 1023–1041.
2001. A mechanism for initiating RNA-dependent RNA polymeriza- Estes, M.K., 2001. Rotaviruses and their replication. In: Knipe, D.M.,
tion. Nature 410, 235–240. Howley, P.M. (Eds.), Fields Virology. Lippincott Willimas & Wilkins,
Caston, J.R., Trus, B.L., Booy, F.P., Wickner, R.B., Wall, J.S., Steven, Philadelphia, pp. 1747–1785.
A.C., 1997. Structure of L-A virus: a specialized compartment for the Estes, M.K., Graham, D.Y., Mason, B.B., 1981. Proteolytic enhancement
transcription and replication of double-stranded RNA. J. Cell Biol. of rotavirus infectivity: molecular mechanisms. J. Virol. 39, 879–888.
138, 975–985. Estes, M.K., Kang, G., Zeng, C.Q., Crawford, S.E., Ciarlet, M., 2001.
Charpilienne, A., Lepault, J., Rey, F., Cohen, J., 2002. Identification of Pathogenesis of rotavirus gastroenteritis. Novartis Found Symp. 238,
rotavirus VP6 residues located at the interface with VP2 that are 82–96 (discussion 96–100).
essential for capsid assembly and transcriptase activity. J. Virol. 76, Fillmore, G.C., Lin, H., Li, J.K., 2002. Localization of the single-stranded
7822–7831. RNA-binding domains of bluetongue virus nonstructural protein NS2.
Chen, D., Luongo, C.L., Nibert, M.L., Patton, J.T., 1999. Rotavirus open J. Virol. 76, 499–506.
cores catalyze 5 -capping and methylation of exogenous RNA: evi- Fiore, L., Greenberg, H.B., Mackow, E.R., 1991. The VP8 fragment of
dence that VP3 is a methyltransferase. Virology 265, 120–130. VP4 is the rhesus rotavirus hemagglutinin. Virology 181, 553–563.
Ciarlet, M., Crawford, S.E., Estes, M.K., 2001. Differential infection Fukuhara, N., Yoshie, O., Kitaoka, S., Konno, T., 1988. Role of VP3 in
of polarized epithelial cell lines by sialic acid-dependent and sialic human rotavirus internalization after target cell attachment via VP7.
acid-independent rotavirus strains. J. Virol. 75, 11834–11850. J. Virol. 62, 2209–2218.
Cohen, J., 2001. Medicine. Rethinking a vaccine’s risk. Science 293, Gajardo, R., Vende, P., Poncet, D., Cohen, J., 1997. Two proline residues
1576–1577. are essential in the calcium-binding activity of rotavirus VP7 outer
Cohen, J., Laporte, J., Charpilienne, A., Scherrer, R., 1979. Activation capsid protein. J. Virol. 71, 2211–2216.
of rotavirus RNA polymerase by calcium chelation. Arch. Virol. 60, Gallegos, C.O., Patton, J.T., 1989. Characterization of rotavirus replication
177–186. intermediates: a model for the assembly of single-shelled particles.
Coulson, B.S., Londrigan, S.L., Lee, D.J., 1997. Rotavirus contains inte- Virology 172, 616–627.
grin ligand sequences and a disintegrin-like domain that are implicated Gillian, A.L., Nibert, M.L., 1998. Amino terminus of reovirus nonstruc-
in virus entry into cells. Proc. Natl. Acad. Sci. U.S.A. 94, 5389–5394. tural protein sigma NS is important for ssRNA binding and nucleo-
Crawford, S.E., Mukherjee, S.K., Estes, M.K., Lawton, J.A., Shaw, A.L., protein complex formation. Virology 240, 1–11.
Ramig, R.F., Prasad, B.V., 2001. Trypsin cleavage stabilizes the ro- Gillian, A.L., Schmechel, S.C., Livny, J., Schiff, L.A., Nibert, M.L., 2000.
tavirus VP4 spike. J. Virol. 75, 6052–6061. Reovirus protein sigmaNS binds in multiple copies to single-stranded
de Haas, F., Paatero, A.O., Mind ich, L., Bamford, D.H., Fuller, S.D., RNA and shares properties with single-stranded DNA binding proteins.
1999. A symmetry mismatch at the site of RNA packaging in the J. Virol. 74, 5939–5948.
polymerase complex of dsRNA bacteriophage phi6. J. Mol. Biol. 294, Gillies, S., Bullivant, S., Bellamy, A.R., 1971. Viral RNA polymerases:
357–372. electron microscopy of reovirus reaction cores. Science 174, 694–696.
Dector, M.A., Romero, P., Lopez, S., Arias, C.F., 2002. Rotavirus gene Gottlieb, P., Strassman, J., Mindich, L., 1992. Protein P4 of the bacte-
silencing by small interfering RNAs. EMBO Rep. 3, 1175–1180. riophage phi 6 procapsid has a nucleoside triphosphate-binding site
Deo, R.C., Groft, C.M., Rajashankar, K.R., Burley, S.K., 2002. Recog- with associated nucleoside triphosphate phosphohydrolase activity. J.
nition of the rotavirus mRNA 3 consensus by an asymmetric NSP3 Virol. 66, 6220–6222.
homodimer. Cell 108, 71–81. Gouet, P., Diprose, J.M., Grimes, J.M., Malby, R., Burroughs, J.N.,
Diprose, J.M., Burroughs, J.N., Sutton, G.C., Goldsmith, A., Gouet, P., Zientara, S., Stuart, D.I., Mertens, P.P., 1999. The highly ordered
Malby, R., Overton, I., Zientara, S., Mertens, P.P., Stuart, D.I., Grimes, double-stranded RNA genome of bluetongue virus revealed by crys-
J.M., 2001. Translocation portals for the substrates and products of a tallography. Cell 97, 481–490.
80 H. Jayaram et al. / Virus Research 101 (2004) 67–81

Grimes, J., Basak, A.K., Roy, P., Stuart, D., 1995. The crystal structure Livolant, F., Levelut, A.M., Doucet, J., Benoit, J.P., 1989. The highly
of bluetongue virus VP7. Nature 373, 167–170. concentrated liquid-crystalline phase of DNA is columnar hexagonal.
Grimes, J.M., Burroughs, J.N., Gouet, P., Diprose, J.M., Malby, R., Zien- Nature 339, 724–726.
tara, S., Mertens, P.P., Stuart, D.I., 1998. The atomic structure of the Lu, G., Zhou, Z.H., Baker, M.L., Jakana, J., Cai, D., Wei, X., Chen, S.,
bluetongue virus core. Nature 395, 470–478. Gu, X., Chiu, W., 1998. Structure of double-shelled rice dwarf virus.
Grimes, J.M., Jakana, J., Ghosh, M., Basak, A.K., Roy, P., Chiu, W., J. Virol. 72, 8541–8549.
Stuart, D.I., Prasad, B.V., 1997. An atomic model of the outer layer Ludert, J.E., Ruiz, M.C., Hidalgo, C., Liprandi, F., 2002. Antibodies
of the bluetongue virus core derived from X-ray crystallography and to rotavirus outer capsid glycoprotein VP7 neutralize infectivity by
electron cryomicroscopy. Structure 5, 885–893. inhibiting virion decapsidation. J. Virol. 76, 6643–6651.
Groft, C.M., Burley, S.K., 2002. Recognition of eIF4G by rotavirus NSP3 Mathieu, M., Petitpas, I., Navaza, J., Lepault, J., Kohli, E., Pothier, P.,
reveals a basis for mRNA circularization. Mol. Cell 9, 1273–1283. Prasad, B.V., Cohen, J., Rey, F.A., 2001. Atomic structure of the
Guerrero, C.A., Mendez, E., Zarate, S., Isa, P., Lopez, S., Arias, C.F., major capsid protein of rotavirus: implications for the architecture of
2000. Integrin alpha(v)beta(3) mediates rotavirus cell entry. Proc. Natl. the virion. EMBO J. 20, 1485–1497.
Acad. Sci. U.S.A. 97, 14644–14649. Mendez, E., Arias, C.F., Lopez, S., 1996. Interactions between the two
Harvey, J.D., Bellamy, A.R., Earnshaw, W.C., Schutt, C., 1981. Biophysi- surface proteins of rotavirus may alter the receptor-binding specificity
cal studies of reovirus type 3. IV. Low-angle X-ray diffraction studies. of the virus. J. Virol. 70, 1218–1222.
Virology 112, 240–249. Mendez, E., Lopez, S., Cuadras, M.A., Romero, P., Arias, C.F., 1999.
Hewat, E.A., Booth, T.F., Loudon, P.T., Roy, P., 1992. Three-dimensional Entry of rotaviruses is a multistep process. Virology 263, 450–459.
reconstruction of baculovirus expressed bluetongue virus core-like Mindich, L., 1988. Bacteriophage phi 6: a unique virus having a
particles by cryo-electron microscopy. Virology 189, 10–20. lipid-containing membrane and a genome composed of three dsRNA
Hewish, M.J., Takada, Y., Coulson, B.S., 2000. Integrins alpha2beta1 and segments. Adv. Virus Res. 35, 137–176.
alpha4beta1 can mediate SA11 rotavirus attachment and entry into Mindich, L., 1999. Precise packaging of the three genomic segments of
cells. J. Virol. 74, 228–236. the double-stranded-RNA bacteriophage phi6. Microbiol. Mol. Biol.
Hill, C.L., Booth, T.F., Prasad, B.V., Grimes, J.M., Mertens, P.P., Sutton, Rev. 63, 149–160.
G.C., Stuart, D.I., 1999. The structure of a cypovirus and the functional Mundt, E., Vakharia, V.N., 1996. Synthetic transcripts of double-stranded
organization of dsRNA viruses. Nat. Struct. Biol. 6, 565–568. Birnavirus genome are infectious. Proc. Natl. Acad. Sci. U.S.A. 93,
11131–11136.
Jayaram, H., Taraporewala, Z., Patton, J.T., Prasad, B.V., 2002. Ro-
Naitow, H., Tang, J., Canady, M., Wickner, R.B., Johnson, J.E., 2002.
tavirus protein involved in genome replication and packaging exhibits
L-A virus at 3.4 Å resolution reveals particle architecture and mRNA
a HIT-like fold. Nature 417, 311–315.
decapping mechanism. Nat. Struct. Biol. 10, 725–728.
Kaljot, K.T., Shaw, R.D., Rubin, D.H., Greenberg, H.B., 1988. Infectious
Nandi, P., Charpilienne, A., Cohen, J., 1992. Interaction of rotavirus
rotavirus enters cells by direct cell membrane penetration, not by
particles with liposomes. J. Virol. 66, 3363–3367.
endocytosis. J. Virol. 62, 1136–1144.
Nason, E.L., Samal, S.K., Venkataram Prasad, B.V., 2000. Trypsin-induced
Kattoura, M.D., Chen, X., Patton, J.T., 1994. The rotavirus RNA-binding
structural transformation in aquareovirus. J. Virol. 74, 6546–6555.
protein NS35 (NSP2) forms 10S multimers and interacts with the viral
Padilla-Noriega, L., Dunn, S.J., Lopez, S., Greenberg, H.B., Arias, C.F.,
RNA polymerase. Virology 202, 803–813.
1995. Identification of two independent neutralization domains on the
Keljo, D.J., Kuhn, M., Smith, A., 1988. Acidification of endosomes is
VP4 trypsin cleavage products VP5, and VP8, of human rotavirus
not important for the entry of rotavirus into the cell. J. Pediatr.
ST3. Virology 206, 148–154.
Gastroenterol. Nutr. 7, 257–263.
Patton, J.T., Spencer, E., 2000. Genome replication and packaging of
Kirkwood, C.D., Bishop, R.F., Coulson, B.S., 1996. Human rotavirus VP4
segmented double-stranded RNA viruses. Virology 277, 217–225.
contains strain-specific, serotype-specific and cross-reactive neutral-
Pesavento, J.B., Lawton, J.A., Estes, M.E., Venkataram Prasad, B.V., 2001.
ization sites. Arch. Virol. 141, 587–600.
The reversible condensation and expansion of the rotavirus genome.
Kobayashi, N., Taniguchi, K., Urasawa, S., 1990. Identification of op- Proc. Natl. Acad. Sci. U.S.A. 98, 1381–1386.
erationally overlapping and independent cross-reactive neutralization Petrie, B.L., Greenberg, H.B., Graham, D.Y., Estes, M.K., 1984. Ultra-
regions on human rotavirus VP4. J. Gen. Virol. 71 (Pt 11), 2615–2623. structural localization of rotavirus antigens using colloidal gold. Virus
Lawton, J.A., Estes, M.K., Prasad, B.V., 1997. Three-dimensional visual- Res. 1, 133–152.
ization of mRNA release from actively transcribing rotavirus particles. Poncet, D., Laurent, S., Cohen, J., 1994. Four nucleotides are the minimal
Nat. Struct. Biol. 4, 118–121. requirement for RNA recognition by rotavirus non-structural protein
Lawton, J.A., Estes, M.K., Prasad, B.V., 1999. Comparative structural anal- NSP3. EMBO J. 13, 4165–4173.
ysis of transcriptionally competent and incompetent rotavirus–antibody Prasad, B.V., Burns, J.W., Marietta, E., Estes, M.K., Chiu, W., 1990. Lo-
complexes. Proc. Natl. Acad. Sci. U.S.A. 96, 5428–5433. calization of VP4 neutralization sites in rotavirus by three-dimensional
Lawton, J.A., Estes, M.K., Prasad, B.V., 2000. Mechanism of genome cryo-electron microscopy. Nature 343, 476–479.
transcription in segmented dsRNA viruses. Adv. Virus Res. 55, 185– Prasad, B.V., Rothnagel, R., Zeng, C.Q., Jakana, J., Lawton, J.A., Chiu,
229. W., Estes, M.K., 1996. Visualization of ordered genomic RNA and
Liemann, S., Chandran, K., Baker, T.S., Nibert, M.L., Harrison, S.C., localization of transcriptional complexes in rotavirus. Nature 382,
2002. Structure of the reovirus membrane-penetration protein, Mu1, 471–473.
in a complex with is protector protein, Sigma3. Cell 108, 283–295. Prasad, B.V., Wang, G.J., Clerx, J.P., Chiu, W., 1988. Three-dimensional
Lima, C.D., Klein, M.G., Hendrickson, W.A., 1997. Structure-based anal- structure of rotavirus. J. Mol. Biol. 199, 269–275.
ysis of catalysis and substrate definition in the HIT protein family. Prasad, B.V.V., Estes, M.K., 2000. Electron cryomicroscopy and com-
Science 278, 286–290. puter image processing techniques: use in structure–function studies
Lima, C.D., Klein, M.G., Weinstein, I.B., Hendrickson, W.A., 1996. of rotavirus. Human Press, New Jersey.
Three-dimensional structure of human protein kinase C interacting Prasad, B.V.V., Yamaguchi, S., Roy, P., 1992. Three-dimensional structure
protein 1, a member of the HIT family of proteins. Proc. Natl. Acad. of single-shelled bluetongue virus. J. Virol. 66, 2135–2142.
Sci. U.S.A. 93, 5357–5362. Reinisch, K.M., 2002. The dsRNA Viridae and their catalytic capsids.
Liu, M., Mattion, N.M., Estes, M.K., 1992. Rotavirus VP3 expressed in Nat. Struct. Biol. 9, 714–716.
insect cells possesses guanylyltransferase activity. Virology 188, 77– Reinisch, K.M., Nibert, M.L., Harrison, S.C., 2000. Structure of the
84. reovirus core at 3.6 Å resolution. Nature 404, 960–967.
H. Jayaram et al. / Virus Research 101 (2004) 67–81 81

Ruiz, M.C., Alonso-Torre, S.R., Charpilienne, A., Vasseur, M., Michelan- inhibition of the transcriptase activity of the rotavirus DLP: a structural
geli, F., Cohen, J., Alvarado, F., 1994. Rotavirus interaction with iso- view. J. Mol. Biol. 307, 161–172.
lated membrane vesicles. J. Virol. 68, 4009–4016. Tihova, M., Dryden, K.A., Bellamy, A.R., Greenberg, H.B., Yeager, M.,
Ruiz, M.C., Charpilienne, A., Liprandi, F., Gajardo, R., Michelangeli, 2001. Localization of membrane permeabilization and receptor binding
F., Cohen, J., 1996. The concentration of Ca2+ that solubilizes outer sites on the VP4 hemagglutinin of rotavirus: implications for cell
capsid proteins from rotavirus particles is dependent on the strain. J. entry. J. Mol. Biol. 314, 985–992.
Virol. 70, 4877–4883. Valenzuela, S., Pizarro, J., Sandino, A.M., Vasquez, M., Fernandez, J.,
Sabara, M., Gilchrist, J.E., Hudson, G.R., Babiuk, L.A., 1985. Prelimi- Hernandez, O., Patton, J., Spencer, E., 1991. Photoaffinity labeling
nary characterization of an epitope involved in neutralization and cell of rotavirus VP1 with 8-azido-ATP: identification of the viral RNA
attachment that is located on the major bovine rotavirus glycoprotein. polymerase. J. Virol. 65, 3964–3967.
J. Virol. 53, 58–66. van Regenmortel, M.H.V., Fauquet, C.M., Bishop, D.H.L., Carstens, E.B.,
Schoehn, G., Moss, S.R., Nuttall, P.A., Hewat, E.A., 1997. Structure of Estes, M.K., Lemon, S.M., Maniloff, J., Mayo, M.A., McGeoch, D.J.,
Broadhaven virus by cryoelectron microscopy: correlation of structural Pringle, C.H., Wickner, R.B., 2000. Virus Taxonomy—the Classifica-
and antigenic properties of Broadhaven virus and bluetongue virus tion and Nomenclature of Viruses, Academic Press, San Diego.
outer capsid proteins. Virology 235, 191–200. Wentz, M.J., Patton, J.T., Ramig, R.F., 1996. The 3 -terminal consensus se-
Shaw, A.L., Rothnagel, R., Chen, D., Ramig, R.F., Chiu, W., Prasad, B.V., quence of rotavirus mRNA is the minimal promoter of negative-strand
1993. Three-dimensional visualization of the rotavirus hemagglutinin RNA synthesis. J. Virol. 70, 7833–7841.
structure. Cell 74, 693–701. Wood, H.A., Bozarth, R.F., 1972. Properties of viruslike particles of
Shaw, A.L., Samal, S.K., Subramanian, K., Prasad, B.V., 1996. The Penicillium chrysogenum: one double-stranded RNA molecule per
structure of aquareovirus shows how the different geometries of the two particle. Virology 47, 604–609.
layers of the capsid are reconciled to provide symmetrical interactions Xia, Q., Jakana, J., Zhang, J.Q., Zhou, Z.H., 2003. Structural comparisons
and stabilization. Structure 4, 957–967. of empty and full cytoplasmic polyhedrosis virus. Protein–RNA in-
Skehel, J.J., Joklik, W.K., 1969. Studies on the in vitro transcription of teractions and implications for endogenous RNA transcription mech-
reovirus RNA catalyzed by reovirus cores. Virology 39, 822–831. anism. J. Biol. Chem. 278, 1094–1100.
Smith, R.E., Furuichi, Y., 1982. The double-stranded RNA genome seg- Yeager, M., Berriman, J.A., Baker, T.S., Bellamy, A.R., 1994.
ments of cytoplasmic polyhedrosis virus are independently transcribed. Three-dimensional structure of the rotavirus haemagglutinin VP4 by
J. Virol. 41, 326–329. cryo-electron microscopy and difference map analysis. EMBO J. 13,
Tao, Y., Farsetta, D.L., Nibert, M.L., Harrison, S.C., 2002. RNA synthesis 1011–1018.
in a cage-structural studies of reovirus polymerase lambda3. Cell 111, Yeager, M., Dryden, K.A., Olson, N.H., Greenberg, H.B., Baker, T.S.,
733–745. 1990. Three-dimensional structure of rhesus rotavirus by cryoelectron
Taraporewala, Z., Chen, D., Patton, J.T., 1999. Multimers formed by the microscopy and image reconstruction. J. Cell Biol. 110, 2133–2144.
rotavirus nonstructural protein NSP2 bind to RNA and have nucleoside Yu, X., Hingorani, M.M., Patel, S.S., Egelman, E.H., 1996. DNA is bound
triphosphatase activity. J. Virol. 73, 9934–9943. within the central hole to one or two of the six subunits of the T7
Taraporewala, Z.F., Chen, D., Patton, J.T., 2001. Multimers of the blue- DNA helicase. Nat. Struct. Biol. 3, 740–743.
tongue virus nonstructural protein, NS2, possess nucleotidyl phos- Zarate, S., Espinosa, R., Romero, P., Guerrero, C.A., Arias, C.F., Lopez,
phatase activity: similarities between NS2 and rotavirus NSP2. Virol- S., 2000. Integrin alpha2beta1 mediates the cell attachment of the
ogy 280, 221–231. rotavirus neuraminidase-resistant variant nar3. Virology 278, 50–
Taraporewala, Z.F., Patton, J.T., 2001. Identification and characterization 54.
of the helix-destabilizing activity of rotavirus nonstructural protein Zhou, Z.H., Baker, M.L., Jiang, W., Dougherty, M., Jakana, J., Dong, G.,
NSP2. J. Virol. 75, 4519–4527. Lu, G., Chiu, W., 2001. Electron cryomicroscopy and bioinformatics
Thouvenin, E., Schoehn, G., Rey, F., Petitpas, I., Mathieu, M., Vaney, suggest protein fold models for rice dwarf virus. Nat. Struct. Biol. 8,
M.C., Cohen, J., Kohli, E., Pothier, P., Hewat, E., 2001. Antibody 868–873.

Potrebbero piacerti anche