You are on page 1of 18

Nidoviruses

Edited by S. Perlman, T. Gallagher, and E. J. Snijder


© 2008 ASM Press, Washington, DC

Chapter 3

Nidovirus Genome Organization and Expression Mechanisms


Paul Britton and Dave Cavanagh

The order Nidovirales comprises three families of replicase proteins, the positive-sense gRNA is copied
viruses with linear single-stranded RNA genomes of into negative-sense counterparts which act as tem-
positive polarity: the Coronaviridae, comprising the plates for the synthesis of new gRNAs. In addition to
two genera Coronavirus and Torovirus, together with a negative-sense gRNA, nidoviruses produce a series
Arteriviridae and Roniviridae. The nidovirus mem- of negative-sense counterparts of the sg mRNAs. The
bers have some common features with respect to synthesis of both full-length and subgenome-length
genome organization, replication, transcription, and negative-sense RNAs is initiated at the 3 end of the
translation of gene products. The name Nidovirales, gRNA. Synthesis of negative-sense RNAs may termi-
derived from the Latin nidus for nest, reflects the nested- nate at different points along the gRNA template,
set arrangement of the subgenomic mRNAs (sg mRNAs) yielding subgenome-length minus-strand RNAs.
produced by this order of viruses during their replica- Attenuation of minus-strand RNA synthesis occurs
tion cycle. The coronaviruses, toroviruses, and roni- at sequences, known as transcription regulatory
viruses possess the longest known RNA genomes, sequences (TRSs), which are well conserved in a virus
ranging from 26 to 32 kb (7, 9, 27, 28, 49, 77, 79), type but differ between groups, genera, and families
and are related by some common features to the arteri- of viruses. Failure to terminate at any of the TRSs
viruses, which have smaller genomes, ranging from 13 results in the synthesis of a full-length negative-sense
to 16 kb (37, 49) (Fig. 1). The coronaviruses are pres- RNA (or “antigenome”). The negative-sense sgRNAs
ently divided into three groups. The groupings were of coronaviruses and arteriviruses acquire a leader
based initially on cross-reactivity in neutralization and sequence of 65 to 100 nucleotides for coronaviruses
immunofluorescence assays but more recently on phy- but 150 to 210 nucleotides for arteriviruses, copied
logenetic relationships and genome organization. The from the 5 end of the gRNA, as part of the process
features of the various coronavirus subgroups are of discontinuous minus-strand RNA synthesis. The
summarized in greater detail in chapters 2, 12, and 15. negative-sense sgRNAs act as templates for the syn-
Nidoviruses replicate in the cytoplasm of an infected thesis of the positive-sense sgRNAs, which are usu-
cell, although one of the structural proteins, the ally generated in a large excess compared to their
nucleocapsid (N) protein, for some viruses has been negative-sense counterparts. The mechanism for the
found to be located within the nucleolar region of the synthesis of nidovirus sgRNAs is called discontinu-
nucleus in virus infected cells (25, 56, 111, 113, 142, ous extension of minus-strand RNA (115–118;
153, 156). For general reviews, please see references reviewed in chapter 8 and reference 119).
Copyright © 2007. ASM Press. All rights reserved.

42, 43, 77, 90, 125, 126, and 137. Toroviruses have two transcription strategies
This chapter covers the genome organization involved in their production of the sg mRNAs. Only the
and expression mechanisms of the nidoviruses as out- largest sg mRNA, RNA2, contains a leader sequence
lined in the five parts numbered in Fig. 2. Following that is derived from the 5 end of the gRNA and is
infection of a susceptible cell by a nidovirus and thought to be added by a mechanism involving discon-
uncoating of the RNA genome, the first step in a suc- tinuous RNA synthesis, according to the current
cessful replication cycle is the production of the repli- model, analogous to the method used for sgRNA syn-
case proteins. The nidovirus genomic RNA (gRNA) thesis for corona- and arteriviruses (148). The other
initially acts as a eukaryotic mRNA for the translation three torovirus sg mRNAs are colinear with the 3 end
of the replicase proteins. Following synthesis of the of the genome and do not contain a 5 leader sequence

Paul Britton and Dave Cavanagh • Division of Microbiology, Institute for Animal Health, Compton, Newbury, United Kingdom.

29
Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
30 BRITTON AND CAVANAGH

Figure 1. Comparison of the genomic organizations of selected nidoviruses. A representative virus from each coronavirus
group is included. The genomes are drawn to scale to emphasize that the first two-thirds of most nidovirus genomes consist of
the replicase genes, ORF1a and ORF1b. The 1 frameshift site (RFS) for each virus is indicated. The drawings are derived
from complete genome sequences of the representative viruses: porcine TGEV, MHV, SARS-CoV, avian IBV, bovine torovirus
(BToV), EAV, and GAV. All the nonreplicase nonstructural protein genes are indicated with open boxes. The structural genes
are those encoding S, M, E, N, HE, and GP proteins; I is an internal ORF identified in the N protein genes of some group 2
coronaviruses. S is shown as a double-shaded box, with the GP116/GP64 region of GAV having some structural similarities to
an S gene. E is indicated as a dark gray box, M as a black box, and N as light gray box. The 5 end of the GAV GP116/GP64
is colored black, as this part of the gene product is predicted to have triple membrane-spanning motifs that are reminiscent of
the M protein. The arterivirus GP5 gene has a function equivalent to that of the S gene in coronaviruses and toroviruses.

derived from the gRNA (128). Analysis of RNA iso- and 12,700 to 15,700 nucleotides for arteriviruses.
lated from the lymphoid organ of prawns (Penaeus The genomes are capped at the 5 end and polyade-
Copyright © 2007. ASM Press. All rights reserved.

monodon) infected with the ronivirus gill-associated nylated at the 3 end (49, 125). In addition, the
virus (GAV), identified two sg mRNAs, neither of genomes contain untranslated regions (UTRs) at their
which contained a leader sequence (29). The nidovirus 5 and 3 termini that play a role in both genome
sg mRNAs, with or without a leader sequence, are replication and synthesis of sgRNAs.
used for the expression of all viral genes with the The overall genomic organization of nidoviruses
exception of the replicase gene products. is 5-UTR-replicase open reading frame 1a (ORF1a)-
replicase ORF1b-nonreplicase genes-3-UTR-poly(A).
However, as illustrated in Fig. 1, apart from the sizes
GENOME ORGANIZATION of the various regions, the major differences between
the nidovirus genomes involve the type and organi-
All nidoviruses have a positive-sense single- zation of the nonreplicase genes, with greater dif-
stranded RNA genome of 26,200 to 31,500 nucleo- ferences occurring between the different families
tides for roniviruses, toroviruses, and coronaviruses of viruses. For example, the coronaviruses and

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 31

Figure 2. Schematic diagram representing the replication cycle of a nidovirus following infection of a susceptible cell. The dia-
gram has five numbered regions that represent the topics discussed in this chapter: (1) gRNA, released from a virus particle
that has infected the cell, to highlight that it initially acts as an mRNA for the translation of the replicase proteins; (2) pro-
grammed 1 frameshifting event, common to all nidoviruses, for the translation of the replicase polyproteins, pp1a and
pp1ab, in differing amounts; (3) another common feature of nidoviruses, proteolytic cleavage of the replicase polyproteins by
virus-encoded proteinases; (4) another feature of nidoviruses, the generation of sg mRNAs (mainly polycistronic but function-
ally monocistronic) for the expression of the other virus-derived proteins; (5) expression strategies used for the translation of
the sg mRNAs into virus proteins. Only the sg mRNAs encoding the structural proteins are shown. The rest of the diagram
represents the interaction of the virus proteins for assembly and release of virus particles. ERGIC, endoplasmic reticulum-
Golgi intermediate compartment.

toroviruses have the general genome organization 5- the features that distinguish this family of viruses
UTR-ORF1a-ORF1b-S-E-M-N-3-UTR-poly(A) (S is from other positive-sense RNA viruses. The two rep-
spike protein, E is envelope protein, and M is integral licase polyproteins, pp1a and pp1ab, are cleaved dur-
membrane protein); it should be noted that no ing de novo synthesis by two types of virus-encoded
sequence for a protein equivalent to the coronavirus E proteinases. There are common motifs within the
protein has been identified in a torovirus genome. As replicase gene that are shared between the different
can be seen from Fig. 1, the organization and type of members of the nidoviruses, characteristic of protein-
genes in the 3-proximal region of arterivirus and ase, polymerase, helicase, and endoribonuclease
ronivirus genomes are very different. activities that are located at specific regions of the
Nidovirus genomes can be divided into two replicase gene. The ORF1b region of the replicase
Copyright © 2007. ASM Press. All rights reserved.

regions, which vary within and between the different encodes two proteins or domains that have not been
virus families within the order Nidovirales. The first identified in other RNA virus families, a zinc-binding
region, gene 1, represents approximately two-thirds domain (ZBD) (27, 50, 124, 130, 146) and a uridylate-
of the genome (20 kb for coronaviruses, toroviruses, specific endoribonuclease (NendoU) (6, 60, 104,
and roniviruses) and encodes the replicase/transcrip- 129), which can be considered genetic markers for
tase functions of the viruses. Two large polyproteins nidoviruses (49).
are expressed from gene 1, polyprotein 1a (pp1a) and
pp1ab; pp1ab is expressed in smaller amounts as a
C-terminal extension of pp1a following a 1 frame- 5 and 3 UTRs
shift event (11–13), as discussed later in this chapter. The 5 UTR is between 200 and 800 nucleotides
The second region of a nidovirus genome encodes the long for coronaviruses and toroviruses, 150 and 220
structural and accessory proteins. The structure and nucleotides for arteriviruses, and 34 nucleotides for
function of the nidovirus replicase genes are two of the ronivirus GAV. The very 5 end of the coronavirus

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
32 BRITTON AND CAVANAGH

and arterivirus UTR is the so-called leader sequence showed a high degree of covariance among different
that is attached to all coronavirus and arterivirus IBV strains, providing phylogenetic support for this
sg mRNAs. The sg mRNAs from roniviruses do not con- structure (136). Similar stem-loop structures were
tain a leader sequence, possibly reflected in the fact predicted in the first 106 nucleotides of the 5 UTR of
that roniviruses have very short 5 UTRs. The 3 UTR the group 2 coronavirus mouse hepatitis virus (MHV),
is between 200 and 500 nucleotides long for coro- although analysis of a further 120 nucleotides identi-
viruses and toroviruses, 60 and 150 nucleotides for fied two more stem-loop structures (151). Analysis of
arteriviruses, and 129 nucleotides for the ronivirus the first 210 nucleotides of the 5 UTR of another
GAV. The nidovirus 3 UTR occurs after the last func- group 2 coronavirus, bovine coronavirus (BCoV),
tional gene, mainly the N protein gene, except for predicted four stem-loop structures, of which stem-
some group 1 coronaviruses, i.e., transmissible gas- loops III and IV showed phylogenetic conservation
troenteritis virus (TGEV), canine coronavirus, and and whose structures were supported by enzymatic
feline coronavirus, which have one or two genes fol- analysis of a BCoV D-RNA (109, 110). Analysis of
lowing the N protein gene, and the roniviruses that the 5 UTR of the arterivirus equine arteritis virus
have a predicted ORF (ORF4) as the last ORF pre- (EAV) has also led to the prediction of secondary
ceding the poly(A) tract (Fig. 1). In addition, some RNA structures, most of which were supported by
coronaviruses that may be similar to infectious bron- phylogenetic conservation and by enzymatic and
chitis virus (IBV) have recently been isolated from a chemical analyses (145).
variety of avian species (geese, mallards, and pigeons) The nidovirus 3 UTRs have also been predicted
that also have one or two ORFs following the N pro- to contain secondary and even tertiary RNA struc-
tein gene (17, 65). The IBV 3 UTR can be divided tures that have been implicated in RNA synthesis
into two regions; the first is immediately downstream during the nidovirus replication cycle. A 68-nucleo-
of the N protein gene, is hypervariable among IBV tide stem-loop was identified immediately down-
strains, and is deleted from some strains (e.g., M41), stream of the MHV N protein gene (59). A similar
whereas the second region is highly conserved (17). structure has been found in the 3 UTR of the group
A number of RNA secondary structures have 3 coronaviruses, IBV (Fig. 3) and turkey coronavirus
been either predicted or identified within the 5 and (TCoV). There is a high degree of covariance among
3 UTRs of various nidoviruses and play a role in different IBV strains, providing phylogenetic support
both genome replication and synthesis of sgRNAs. for this structure (31). Interestingly, in most IBV
Coronavirus replication signals within the 5 and 3 strains the stem-loop structure is found not adjacent
ends of the genomes were initially investigated using to the end of the N protein gene but at the junction of
defective RNA (D-RNA) genomes; the minimal the hypervariable and conserved regions of the IBV 3
lengths of these regions necessary for replication of UTR. However, in IBV strains lacking the hypervari-
D-RNAs in the presence of helper virus are summa- able region, the stem-loop structure is adjacent to the
rized in Table 1 and include both the UTRs and 5 end of the N protein gene. It is possible that
and 3 regions of the genomes containing parts of the the hypervariable region of the IBV 3 UTR may be
replicase and N protein genes, respectively. Three the remnants of ORFs that have been lost during evo-
stem-loop structures were predicted in the first 100 lution of the virus. In addition to the 3 UTR stem-
nucleotides of the 5 UTR of the group 3 corona- loop structures, an RNA pseudoknot structure was
virus IBV, of which the 5-proximal stem-loop (I) identified downstream of the stem-loop structure,
Copyright © 2007. ASM Press. All rights reserved.

Table 1. Minimal lengths of nidovirus terminal 5 and 3 regions involved in replication

Minimum length (nucleotides)


Family Group Virus
5 end 3 end

Coronaviridae Coronavirus
1 TGEV 649 492
2 MHV 466 436
3 IBV 544 338
Torovirus EToV 607 242
Arteriviridae Arterivirus EAV 296 354
a
Roniviridae Okavirus GAV ND ND
a
ND, not determined.

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 33

sheep, and turkeys and of the picornavirus equine


rhinovirus serotype 2 (64). The sequence representing
the s2m stem-loop structure was also identified in the
3 UTR of the coronaviruses isolated from geese,
pigeons, and mallard ducks (65). Remarkably, an
almost identical sequence and predicted s2m struc-
ture was identified in the 3 UTR of SARS-CoV (89).
A 2.7-Å resolution three-dimensional crystal struc-
ture of the SARS-CoV s2m RNA element has been
determined (112). In addition to the RNA structures
identified in the 3 UTRs of coronaviruses, an octa-
nucleotide motif, 5-GGAAGAGC-3, of unknown
function has been identified 70 to 80 nucleotides
proximal to the poly(A) tract in all coronavirus
genomes examined to date.
Analysis of the nidovirus 3 UTR has been mainly
Figure 3. RNA structural elements in the 3 UTR of the avian coro- restricted to coronaviruses. However, a kissing-loop
navirus IBV. The IBV 3 UTR contains three predicted RNA struc- interaction has been proposed for the arterivirus por-
tures, a stem-loop associated with a pseudoknot structure and a cine reproductive and respiratory virus (PRRSV),
second stem-loop structure, s2m. Similar stem-loop and associated
pseudoknot RNA structures have been identified in the 3 UTRs of
between the loops of stem-loop structures in the 3
other coronaviruses. The s2m structure in the IBV 3 UTR also UTR and N protein gene, to be crucial for minus-
occurs in the 3 UTRs of various astroviruses and a human rhino- strand RNA synthesis (149). A stem-loop structure in
virus as well as in 3 UTR of TCoV, another group 3 coronavirus the 3 UTR of the arterivirus simian hemorrhagic
related to IBV. Neither the s2m structure nor its composite fever virus (SHFV) was identified that contains bind-
sequence was identified in the 3 UTR of any other non-group
3 coronavirus, until the isolation of SARS-CoV. The numbers rep-
ing sites for two cellular proteins (88). A stem-loop
resent the IBV genomic nucleotide positions. structure that may act as a recognition sequence for
the initiation of minus-strand RNA synthesis has
been identified at the 3-terminal end of the arterivi-
initially in the 3 UTR of BCoV (152). The authors rus EAV 3 UTR (5). The identification of potential
postulated that the pseudoknot functions in the plus RNA structures involved in RNA synthesis for toro-
strand as a regulatory control element in coronavirus virus or roniviruses has not been published.
RNA replication. Similar structures were predicted to
be present in the 3 UTRs of other coronaviruses. An
example of a coronavirus 3 UTR stem-loop and Replicase Gene Region
pseudoknot structure is shown in Fig. 3. Further The overall organization of the nidovirus repli-
pro
analysis of the MHV stem-loop/pseudoknot struc- case gene motifs is TM1-TM2-3CL -TM3-RFS-
tures indicated that they may be involved in an RNA RdRp-ZBD-HEL-NendoU, in which TM represents
pro
switch governing a transition between different steps hydrophobic transmembrane domains, 3CL is a
of virus replication (47, 58). Analysis of the 3 UTR chymotrypsin-like proteinase (so called because it has
of the severe acute respiratory syndrome coronavirus substrate specificities similar to those of picornavirus
(SARS-CoV) identified both a stem-loop and a pseu- 3C proteinases), RFS is the ribosomal frameshift site
doknot structure that overlapped in a manner similar required for the expression of the ORF1b-encoded
Copyright © 2007. ASM Press. All rights reserved.

to that identified for MHV and BCoV, potentially proteins, RdRp is the RNA-dependent RNA poly-
reflecting some shared common ancestry between merase, and HEL is a helicase (49). The three hydro-
the viruses (47). Further work demonstrated that phobic TMs have been postulated to be involved
the SARS-CoV 3 UTR could functionally replace the in anchoring the replicase proteins to membrane
MHV 3 UTR, whereas 3 UTRs from group 1 and 3 compartments, forming the so-called replication/
coronaviruses could not replace the MHV 3 UTR, transcription complexes. In addition, other motifs,
providing more evidence that SARS-CoV shares some representing other replicase gene functions but more
common ancestry with group 2 coronaviruses (48). A related to RNA processing, have been identified in
second stem-loop structure, s2m (Fig. 3), was identi- the replicase genes of coronaviruses, toroviruses, and
fied in the 3 UTR of all IBV strains to date, with roniviruses, with one (NendoU) also occurring in
remarkable similarity in both sequence and structure arteriviruses (129). Although the in vitro functions of
to stem-loop structures identified in the genomic 3 some of these gene products have been confirmed,
ends of astroviruses isolated from humans, pigs, their roles in replication/transcription have yet to be

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
34 BRITTON AND CAVANAGH

determined. The proteins include a 3-to-5 exonucle- conserved nucleotide sequences, the TRSs, upstream
ase, a 2-O-ribose methyl transferase, and a cyclic of almost every virus gene. The consequence of the
phosphodiesterase (CPD). The CPD motif (ns2; polymerase terminating at TRSs is that the sgRNAs
Fig. 1) was initially identified in the group 2 corona- can be produced in unequal amounts, in effect con-
virus MHV, not as a part of the replicase gene but as trolling the amounts of the gene products. The fac-
ORF2a in sg mRNA2. The CPD motif is present within tors and/or sequences that regulate whether or how
the replicase gene of toroviruses. A domain corre- the polymerase terminates at a TRS have been an
sponding to an ADP-ribose 1-phosphatase (ADRP) area of considerable interest. Sequences flanking the
has been identified in coronaviruses (nsp3) (129) and TRSs have been postulated to play a role in control-
toroviruses (38). The activity of bacterially expressed ling the levels of sgRNA, as the lengths of identical
ADRP domains from SARS-CoV (114), human coro- nucleotides in flanking regions of a TRS and those at
navirus 229E (HCoV-229E [106]), and TGEV (107) the leader junction vary (55). The roles of the flank-
has been verified in vitro, and the structure of the ing regions have been investigated for both coronavi-
SARS-CoV ADRP has been solved (114). The func- ruses (2, 55, 133, 158) and arteriviruses (35, 98–100,
tion of the nidovirus replicase proteins is discussed in 147). These are discussed in chapter 8.
more detail in chapters 5 and 6. All of the sg mRNAs, except the smallest RNA,
The replicase polyproteins, pp1a and pp1ab, are are structurally polycistronic, but they are usually
proteolytically processed during translation by two functionally monocistronic. The coronaviruses syn-
types of virus-encoded proteinases, a papain-like pro- thesize between five (IBV) and eight (SARS-CoV)
pro
teinase (PL ) and a chymotrypsin-like viral main sg mRNAs encoding four structural proteins, S, E, M,
pro
proteinase, 3CL , into 16 nonstructural proteins and N proteins, and various numbers of nonstruc-
(nsp1 to nsp16) for most coronaviruses, except the tural accessory proteins (Fig. 1). The overall gene
avian coronavirus IBV, which lacks nsp1 and encodes arrangement of the coronavirus structural proteins is
15 (nsp2 to nsp16) replicase gene-derived proteins. highly conserved, following the pattern hemaggluti-
pro
Most coronavirus genomes contain two PL nin-esterase (HE)-S-E-M-N (Fig. 1). Some group 2
domains; however, IBV and SARS-CoV only encode coronaviruses, BCoV, HCoV-OC43, and MHV,
pro
one copy of PL . The arterivirus replicase polypro- encode a fifth structural protein, HE (67). Interest-
teins are cleaved by similar enzymes and may encode ingly, it has been demonstrated that rearrangement of
pro
up to three PL domains, resulting in 13 or 14 prod- this gene order is not deleterious to MHV replication
ucts (nsp1 to nsp12), including the products of a (34). Coronaviruses differ with respect to both the
recently described internal cleavage of nsp7 (144). number and location of the accessory protein genes
Expression of the torovirus replicase gene has been (Fig. 1; reviewed in reference 77) and, in some cases,
postulated to result in 12 proteins (128). The ronivi- in the mode of translation of the proteins, which is
rus replicase gene appears to encode only one pro- discussed later. The toroviruses synthesize four
pro
teinase, equivalent to 3CL , and potentially gives sg mRNAs and have the gene order S-M-HE-N (Fig. 1),
rise to at least four proteins (27, 49). where the HE gene is intact in bovine torovirus but a
pseudogene in equine torovirus (37). The arterivi-
ruses EAV, PRRSV, and lactate dehydrogenase-elevat-
Structural and Accessory Protein Gene Regions ing virus (LDV) synthesize six sg mRNAs all encoding
In contrast to the first (replicase) region of the structural proteins, all of which, apart from the N
genome, the structural and accessory protein gene protein, are M proteins (Fig. 1). The arteriviruses
products are expressed from a series of sg mRNAs. The express proteins equivalent to the coronavirus E, M,
Copyright © 2007. ASM Press. All rights reserved.

sg mRNAs are produced as a “nested” set of RNAs and N proteins, with either four extra membrane gly-
that have different 5 end regions but common 3 coproteins for EAV/PRRSV/LDV and possibly up to
extensions, often referred to as a 3-coterminal nested seven extra M proteins for SHFV, of which three may
set of sg mRNAs; synthesis of nidovirus sgRNAs is have arisen by a putative triple gene duplication
discussed in chapter 8. The sg mRNAs encode the event. Only two viruses, GAV and yellow head virus,
virion structural proteins and, though only for coro- which infect prawns, have been assigned to the family
naviruses, a set of small nonstructural proteins, of Roniviridae, and on the basis of sequence analysis
unknown function, that are referred to as accessory their replicase genes have been shown to be related to
proteins. Synthesis of the coronavirus and arterivirus those of coronaviruses, toroviruses, and arteriviruses.
sgRNAs is via a discontinuous process, as indicated The roniviruses produce two sg mRNAs; both the gene
in the introduction. Negative-sense counterparts of order and the format of the structural proteins are
the sg mRNAs are synthesized from the full-length very different from those identified in the genomes of
gRNA as a result of being attenuated or terminated at coronaviruses, toroviruses, and arteriviruses. A gene

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 35

equivalent to the N protein gene of other nidovi- is expressed and the above observations, ribosomal
ruses is found immediately downstream of ORF1b frameshifting, as a mechanism for translating ORF1b
(Fig. 1). of the nidovirus replicase genes, was confirmed for
In addition to the structural protein genes, the IBV (10–12).
coronavirus genomes contain a variety of accessory The potential for ribosomal frameshifting within
protein genes, which are interspersed among the the IBV replicase gene was shown by the insertion of
structural protein genes and which in some group 1 the IBV ORF1a/1b overlap region into a reporter
and 3 coronaviruses are found downstream of the N gene and testing of the construct for any frameshift-
protein gene (Fig. 1). A number of laboratories have ing ability in vitro using a transcription and transla-
shown that the accessory protein genes are not essen- tion system (11). Results indicated that the region
tial for virus replication in vitro, in vivo, or ex vivo mediated a 1 frameshift event, with about 30% of
(21, 30, 33, 51, 57, 97, 132). the ribosomes changing frame within the ORF1a/1b
overlap region. Subsequently, similar experiments
were used to demonstrate ORF1a/1b frameshifting
RIBOSOMAL FRAMESHIFTING for other coronaviruses (8, 15, 41, 54, 140), a toro-
virus (equine torovirus) (130), an arterivirus (EAV)
One of the main characteristic features of nido- (36), and a ronivirus (GAV) (27), with putative frame-
virus genomes is the large replicase gene region, of shift sites identified in other sequenced nidovirus
which approximately two-thirds of the genome genomes.
encodes the proteins involved in replication and syn- Following the demonstration that frameshifting
thesis of sgRNAs. The replicase polyproteins, pp1a could occur in IBV, potential signals involved in
and pp1ab, are produced from the same initiation directing the frameshifting events were investigated
codon; pp1ab is a C-terminally extended form of (12, 14, 16, 134). Potential signals were located
pp1a as a result of a programmed 1 ribosomal within the ORF1a/1b junction (Fig. 4) and comprised
frameshifting event occurring in the region where two cis-acting RNA elements, the first representing
ORF1a and ORF1b briefly overlap. As a result, the the heptameric slippery sequence, UUUAAAC, con-
ORF1b-derived replicase proteins are produced in forming to the predicted slippery sequence motif
smaller amounts (approximately 60 to 80% less) XXXYYYN, the site of the actual 1 frameshift
than the ORF1a-derived replicase proteins. The (Fig. 4). In vitro experiments showed that 30% of the
ORF1b-encoded proteins include proteins with key ribosomes that entered the sequence in the ORF1a
replicative functions, like RdRp and HEL activities, frame, U-UUA-AAC, left in the ORF1b frame repre-
indicating that the conserved frameshift mechanism senting the codons UUU-AAA-C (Fig. 5). Interest-
is a way of down-regulating the expression of these ingly, all the coronavirus, torovirus, and arterivirus
enzymes, which must be an important control system genomes sequenced to date have the same UUUAAAC
in the nidovirus life cycle. Programmed 1 ribosomal slippery sequence, apart from EAV, which has the
frameshifting has been reported to occur during the sequence GUUAAAC (Fig. 6). The two ronivirus rep-
translation of other viral genes, including those in the licase sequences available, those for GAV and yellow
animal virus families Retroviridae (61), Astroviridae head virus, have AAAUUUU as their slippery sequence.
(63), Totiviridae (150), and Flaviviridae (26, 154) Experiments showed that the slippery sequence was
and various plant viruses (68, 105). not in itself sufficient to result in the ribosomes’
Completion of the first coronavirus genome changing frame. A second element required for frame-
sequence, for IBV (7), raised the question of the shifting to occur was identified as an RNA pseu-
Copyright © 2007. ASM Press. All rights reserved.

mechanism for expression of ORF1b. Sequence anal- doknot (Fig. 4 and 7), an RNA structure composed of
ysis showed that ORF1a terminated at genome nucle- two base-paired regions stacked coaxially and con-
otide 12382 and that ORF1b, which overlapped nected by two single-stranded loop regions (19, 103).
ORF1a by 42 nucleotides, was in a 1 reading frame No particular nucleotide sequences were identified as
with respect to ORF1a (Fig. 4). No sg mRNA equiva- playing an essential role, although the overall shape
lent to a size required for the expression of ORF1b and predicted stabilities of the structure had to be
had been detected in IBV-infected cells, and no IBV maintained for efficient frameshifting. The primary
TRS, CUUAACAA, had been identified upstream of role of the RNA pseudoknot is believed to involve
ORF1b. Nucleotide comparisons of the IBV slowing down or stalling of the ribosome as it trans-
ORF1a/1b junction with the gag-pol overlap region lates through the slippery sequence (Fig. 7), to allow
of Rous sarcoma virus, a region suspected to elicit a realignment of the decoding tRNA on the mRNA in a
frameshifting event (62), showed significant sequence new frame (Fig. 5). The RNA pseudoknot has to be
identity. As a result of the conundrum of how ORF1b positioned within 5 to 7 nucleotides downstream of

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
36 BRITTON AND CAVANAGH

Figure 4. Sequence of the IBV ribosome frameshifting site. The top drawing represents the IBV gRNA, as shown in Fig. 1
and 3. The sequence representing the junction of ORF1a and ORF1b, corresponding to the RFS site, is expanded below the
gRNA drawing. The numbers represent the IBV genomic nucleotide positions. The positions of the heptameric UUUAAC slip
site sequences (in bold) are shown; the nucleotides forming the stem and loop structures of the pseudoknot (shown in Fig. 6)
are underlined. The amino acid sequences corresponding to ORF1a, -1b, and -1ab are shown below the IBV genomic nucleo-
tide sequence. ORF1a terminates at nucleotide 12382. There is no initiation codon for ORF1b, but a contiguous amino acid-
encoding sequence starts at nucleotide 12342 and terminates at nucleotide 20417. The 1 frameshift site, highlighted in bold,
takes place within the coding context of the slip site in which an asparagine residue, encoded by ORF1a, is replaced by a lysine
residue, encoded by ORF1b, due to the ribosome moving the RNA back one nucleotide (shown in Fig. 5). As a result of the
1 frameshift event, ORF1a is extended by the ORF1b coding sequence and terminates as an ORF1ab fusion protein at
nucleotide 20415. Translation termination codons are marked with asterisks.

the slip site (Fig. 7) for efficient frameshifting (14). the tRNA accommodation step of elongation by
The actual mechanism of frameshifting is thought to filling the entrance of the ribosomal mRNA tunnel
occur by the simultaneous slippage, into the 1 (40, 102). The restriction on the movement of the
frame, of two ribosome-bound tRNAs within the RNA can be eased in two ways: (i) unwinding or melt-
aminoacyl and peptidyl sites of the ribosome during ing of the pseudoknot, which allows the RNA to move
the translation of the slippery sequence (44, 52, 86). For forward, or (ii) movement or slippage of the RNA one
Copyright © 2007. ASM Press. All rights reserved.

the coronavirus IBV the incorporation of the amino nucleotide backwards, resulting in the 1 frameshift
acid lysine, derived from the ORF1b sequence, into (reviewed in reference 52). Cryoelectron microscopy
pp1a, instead of the amino acid asparagine (Fig. 5), of mammalian 80S ribosomes paused at a coronavirus
results in the generation of pp1ab. pseudoknot revealed an intermediate of frameshifting
Although the pseudoknot structures cause the in which it was possible to determine how the pseu-
ribosome to pause, resulting in the ribosomal amino- doknot interacts with the ribosome (96).
acyl and peptidyl sites being placed over the slippery
sequence (70), the pausing is insufficient for 1 frame-
shifting to occur (143); the duration of the pausing REGULATION AND EXPRESSION STRATEGIES
does not necessarily correlate with the level of frame-
shifting observed (70, 101). Crystallographic studies, The nidovirus gRNAs, like the genomes of other
in concert with molecular studies, indicate that the positive-strand RNA viruses, such as the Flaviviridae
7
pseudoknot restricts movement of the RNA during and Togaviridae, contain an m GpppN cap structure

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 37

Figure 5. Schematic diagram representing the IBV 1 frameshift event for the synthesis of pp1ab. The top part shows the
progression of a ribosome along the IBV gRNA over the UUUAAAC slip site. Shown are the positions of the aminoacyl-tRNAs
decoding the ORF1a codons, elongation of the polypeptide chain, and decoding of the next codon, resulting in the synthesis
of pp1a. The lower part represents a 1 frameshift event as proposed by the simultaneous-slippage model, in which the
ribosome-bound aminoacyl-tRNAs are proposed to slip simultaneously one nucleotide to a 1 frame position from their ini-
tial frame. Frameshifting can only occur when the anticodons of the two aminoacyl-tRNAs associated with the ribosome and
mRNA can still form two base pairs with the RNA in the shifted 1 frame, indicated by the first and second anticodon-codon
Asn
base pairings, but with disruption of base pairing at the third position. Following the 1 slippage event, aminoacyl-tRNA
Lys
dissociates from the ribosome complex, before decoding the mRNA, allowing the next aminoacyl-tRNA, aminoacyl-tRNA ,
to move into the aminoacyl site of the ribosome, in which there is full complementarity between the anticodon of the
tRNA and codon on the mRNA in the ORF1b frame. The mRNA is then decoded at this position, allowing elongation and
movement of the ribosome to the next codon. The IBV replicase gene is now decoded in the ORF1b frame, resulting in transla-
tion of pp1ab.

at the 5 end. This essentially makes the gRNAs coronaviruses HCoV-229E and MHV (141), although
into mRNAs, imparting a dual role in which the the presence of N protein was later shown to increase
gRNAs can act as a template for the synthesis of the efficiency of rescue and replication of HCoV-229E
negative-sense RNAs (replication) and for the trans- (120) and TGEV (1) replicons. This indicates that the
lation of the replicase polyproteins. As nidovirus N protein has some effect on coronavirus replication,
gRNAs and sg mRNAs contain a 5 cap structure, possibly by protecting the viral RNAs, or involve-
ribosomes are able to initiate translation in a cap- ment in translation of the viral RNAs; however, N
Copyright © 2007. ASM Press. All rights reserved.

dependent manner. protein is dispensable for synthesis of arterivirus


There is evidence from the analysis of some coro- RNA (95).
naviruses that translation can be regulated by both Nidovirus RNAs are very similar to host cell
viral and host cell factors acting in trans or by cis- mRNAs and therefore dependent on the availability
acting sequences within the virus 5 UTR (122). An and activities of host cellular factors, although, as
interaction of the MHV N protein with the leader indicated above, various viral RNA elements or virus-
sequence has been proposed to stimulate translation encoded products (N protein) may increase the effi-
of viral mRNAs (138, 139). Interestingly, the IBV N ciency of translation. Cellular initiation factors bind
7
protein was found to be essential for the rescue of to the m GpppN cap structure at the 5 end of the
IBV from intracellular T7 polymerase-generated viral RNAs for recruitment of the 43S ribosomal
infectious RNA in IBV reverse genetics systems (22, complex and formation of the 48S ribosomal com-
157). N protein was not required for the rescue of plex, which scans the mRNA for AUG initiation
infectious viruses in reverse genetics systems for the codons. On encountering the AUG codon, GTP is

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
38 BRITTON AND CAVANAGH
Copyright © 2007. ASM Press. All rights reserved.

Figure 6. Comparison of the nidovirus replicase gene frameshifting sites. (A) Independent alignment of the gRNA sequences,
from the slip site, over the pseudoknot sequences of various coronavirus, torovirus, arterivirus, and ronivirus sequences.
Nucleotides common within the grouped sequences are highlighted in black. The RFS slip site is underlined. (B) Phylogenetic
groupings of all the aligned nidovirus sequences from panel A to demonstrate that as well as falling within their genus group-
ings, the coronavirus-derived sequences also fall within their groups, indicating that the viruses are related through their RFSs.
FIPV, feline infectious peritonitis virus; PEDV, porcine epidemic diarrhea virus; BToV, bovine torovirus; YHV, yellow head
virus.

hydrolyzed, initiation factors are released, and the involved in binding of the initiation complex, can be
60S ribosomal subunit binds to the preinitiation com- phosphorylated, which increases its binding to the
7
plex for formation of the 80S ribosome for transla- m GpppN cap structure at the 5 end of eukaryotic
tion and elongation of the virus product (reviewed in mRNAs and to eIF-G4, usually enhancing translation
reference 39). One initiation factor, eukaryotic initia- rates (46). The phosphorylation of eIF-4E plays a
tion factor 4E (eIF-4E), the cap-binding protein regulatory role in translation initiation, and viruses

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 39

Figure 7. Schematic diagram indicating the pausing of ribosomes by a pseudoknot structure. The top part represents ribosomes
encountering a stem-loop structure on an mRNA being decoded. The ribosomes may be slowed down by such a structure, but
they are able to melt the structure and decode the mRNA. The lower part represents ribosomes encountering a pseudoknot
structure with an upstream slip site. Provided the slip site and pseudoknot are separated by no more than 5 to 7 nucleotides,
the pausing effect of the pseudoknot can cause a 1 frameshift as illustrated in Fig. 6. The predicted RNA structure of the IBV
pseudoknot is shown as an example to highlight the nucleotides forming the slip site and the stem and loop structures.

have utilized this process for controlling host protein are required for synthesis of negative-strand RNA
synthesis (reviewed in reference 93). Increased phos- (82). Evidence for interactions between the 5 and 3
phorylation of eIF-4E has been observed in cells ends of nidovirus genomes has not been documented.
infected with the coronaviruses MHV (4) and SARS- As indicated in the introduction, the 5 UTRs of the
CoV (92) as a result of the activation of p38 mitogen- gRNA are between 200 and 800 nucleotides long for
activated protein kinase. coronaviruses and toroviruses and 150 and 220 nucle-
otides for arteriviruses. Although the 5 UTRs of coro-
naviruses and arteriviruses are relatively devoid of
Regulation and Expression of the Replicase Gene initiation codons, a small ORF, the intraleader ORF,
A complexity for the translation of RNAs from coding for 3 to 11 amino acids, is found in the corona-
viruses with a positive-sense RNA genome is associ- virus 5 UTRs. These include ORFs of 24 nucleotides,
ated with the gRNA, which can act either as an mRNA beginning at position 99 in the 209-nucleotide 5 UTR
or as a template for replication. The signals for replica- of MHV-A59 (80); 24 nucleotides, beginning at posi-
tion can be in either the 5 or 3 UTR or both. Interac- tion 101 in the 210-nucleotide 5 UTR of BCoV (109);
tions between the 5 and 3 ends of the gRNA of some 33 nucleotides, beginning at position 81 in the 293-
positive-sense RNA viruses have been shown to con- nucleotide 5 UTR of HCoV-229E (53); 9 nucleotides,
trol the switch between translation and replication beginning at position 117 in the 313-nucleotide 5
(39). As discussed above, the 5 and 3 UTRs of the UTR of TGEV (41); and 33 nucleotides, beginning at
Copyright © 2007. ASM Press. All rights reserved.

nidovirus genomes contain a variety of RNA struc- position 131 in the 528-nucleotide 5 UTR of IBV (7).
tures that may be involved in interactions with viral A slightly larger intraleader ORF, of 111 nucleotides,
and cellular proteins. Most studies have involved inter- is found beginning at position 14 in the 211-nucleotide
actions of proteins with the 3 UTR of group 2 coro- 5 UTR of the arterivirus EAV (66). The intraleader
naviruses. For example, the poly(A) tail of MHV and ORF initiation codon was identified as forming part of
BCoV was found to be an important cis-acting signal the stem structure of stem-loop III of the group 2 coro-
for replication and was found to bind poly(A)-binding navirus BCoV. The authors indicated that such ORFs
protein (135). Studies using MHV D-RNAs indicated may be part of stem-loop structures identified in the 5
that 466 nucleotides from the 5 end and 436 nucleo- UTRs of other coronaviruses (109). The identification
tides from the 3 end of the genome are required for of such ORFs led to the proposal that they may play a
D-RNA replication (69, 81) (Table 1). In addition, role in regulation of translation of the replicase gene
both ends of the MHV genome are required for posi- from the gRNA (122). Recombinant EAVs in which
tive-strand synthesis, but only the last 55 nucleotides either the Kozak sequence of the intraleader ORF was

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
40 BRITTON AND CAVANAGH

optimized or the initiation codon was incapacitated example, TGEV, in which the 3a and 3b genes are
exhibited similar phenotypes but showed reduced rep- expressed from sg mRNA3 (18), and HCoV-229E, in
lication kinetics, plaque size, and viral yields compared which the 4a and 4b genes are expressed from
to those of the parental virus (3). It should be noted sg mRNA4 (108). Other group 1 coronaviruses have
that inactivation of the EAV intraleader ORF by Molen- more than one functionally bi- or tricistronic sg mRNA,
kamp et al. (94) did not have an effect on virus growth, for example, feline coronavirus, in which sg mRNA3
indicating that the results obtained by Archambault et has the capacity to express three genes (3a, 3b, and 3c
al. (3) could be due to changes in RNA structure rather [51, 155]) and sg mRNA7 produces two genes (7a and
than intraleader ORF expression. 7b [32]). In general, the group 2 coronaviruses have
only one functionally bicistronic sg mRNA, e.g.,
sg mRNA5 of MHV, which expresses two genes, 5a
Regulation and Expression of Genes Encoding and 5b, the latter encoding the E protein (20, 127).
Structural and Accessory Proteins SARS-CoV yields three functionally bicistronic
The nidovirus nonreplicase proteins, the struc- sg mRNAs, sg mRNA3, -7, and -8, which produce the
tural and accessory proteins, are not produced as poly- 3a and 3b, 7a and 7b, and 8a and 8b proteins, respec-
proteins but are expressed from the sg mRNAs. In most tively (140). In all the cases described above, the
cases the cap-dependent mechanism directs translation functionally bi- or tricistronic sg mRNAs express two
of the 5-most ORF from an sg mRNA, which apart or three ORFs from the 5 end of the mRNA. However,
from the smallest species are structurally polycistronic some type 2 coronaviruses, MHV (45), BCoV (123),
but functionally monocistronic. The synthesis of the sg and SARS-CoV (140), have a different type of func-
mRNAs is not equimolar, although in most cases the tionally bicistronic sg mRNA to express an I ORF
smallest sg mRNA, usually producing the N protein, is within the N protein gene. The I ORF is in a 1 read-
produced in the largest amounts. Although the group 1 ing frame located downstream of the N protein gene
coronaviruses and some newly identified group 3 coro- initiation codon. Recent studies of the structure of the
naviruses have ORFs downstream of the N protein SARS-CoV I ORF (ORF9b) indicated that the protein
sequence, these smaller sg mRNAs are often produced has a dimeric tent-like  structure with an amphipa-
in much smaller amounts than other sg mRNAs. In thic surface and a central hydrophobic cavity that
effect, a control on the amount of sg mRNA determines binds lipid molecules (91). The authors indicated that
the amount of protein produced. Nevertheless, some the hydrophobic cavity was likely to be involved in
coronaviruses have more than one ORF translated membrane attachment and showed the ORF9b protein
from a particular sg mRNA. The majority of these are to associate with intracellular vesicles.
bicistronic, though the group 3 coronavirus IBV and The group 3 coronaviruses IBV, TCoV, and
the related TCoV and pheasant coronavirus (PhCoV) PhCoV produce a functionally tricistronic sg mRNA3,
express a tricistronic sg mRNA (23, 24, 83, 85). which encodes the 3a, 3b, and 3c proteins, where the
All coronaviruses identified to date contain at ORF3c product is the E protein (23, 24, 83, 85), and
least one functionally bicistronic sg mRNA, some of a bicistronic sg mRNA5, which produces the 5a and
which are summarized in Table 2. The number and 5b proteins (23, 24, 84). Arteriviruses also contain a
position of the genes expressed from such sg mRNAs functional bicistronic sg mRNA; the EAV E and gly-
vary with the different types and groups of coronavi- coprotein 2 (GP2) structural proteins are encoded
ruses. The group 1 coronaviruses include viruses that by two overlapping ORFs, ORF2a and ORF2b,
have only one functionally bicistronic sg mRNA, for which are both expressed from sg mRNA2 (131).
Copyright © 2007. ASM Press. All rights reserved.

Table 2. Coronavirus sg mRNAs expressing more than one product


a
Coronavirus Group No. sg mRNA Products

TGEV 1 1 3 3a, 3b
HCoV-229E 1 1 4 4a, 4b
FCoV 1 2 3 and 7 3a, 3b, 3c, 7a, 7b
MHV 2a 2 5 and 8 5a, 5b (E), N, I
SARS-CoV 2b 4 3, 7, 8, and 9 3a, 3b, 7a, 7b, 8a, 8b, N, I
IBV 3 2 3 and 5 3a, 3b, 3c (E), 5a, 5b
TCoV 3 2 3 and 5 3a, 3b, 3c (E), 5a, 5b
PhCoV 3 2 3 and 5 3a, 3b, 3c (E), 5a, 5b
a
It should be noted that the nidovirus gRNA is a functionally bicistronic mRNA for the replicase gene (gene 1), but
structurally a polycistronic mRNA.

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 41

Homologous genes in other arteriviruses, PRRSV, of IBV in cell culture (57). A series of experiments
LDV, and SHFV, are also postulated to be expressed using in vitro-generated RNA transcripts modifying
from functionally bicistronic sg mRNAs. either the Kozak context sequences for the three gene
The second and third ORFs expressed from the 3 ORFs, expression in the absence of the 5 cap ana-
functionally bicistronic and tricistronic sg mRNAs are log 7-methyl-GTP, introduction of 5 structures to
expressed as independent proteins. The simplest mech- prevent expression of ORF3a and ORF3b, or expres-
anism for the expression of a downstream ORF is sion of the three ORFs downstream of the influenza
leaky scanning (71, 72, 74). The ribosome-scanning A virus N protein gene showed that ORF3c was still
model predicts that the efficiency with which a partic- expressed when the expression of ORF3a and ORF3b
ular ORF is translated depends on its location in the had been prevented (85). These observations led the
mRNA and the sequence context of the AUG initia- authors to propose that the expression of 3c was
tion codon (71, 73, 75, 76). In this mechanism, pre- being directed by a cap-independent method involv-
initiation complexes, on encountering the first AUG ing internal initiation that is dependent on the 3ab
initiation codon, either form an active 80S ribosome sequence and that the 3b protein was expressed as the
for de novo synthesis of the associated protein or result of leaky scanning using cap-dependent transla-
bypass the first AUG codon and continue to scan down tion. An RNA structure was predicted and proposed
the mRNA until they encounter another AUG codon. to play a role in the internal entry of ribosomes for
For leaky scanning to occur, the first AUG codon ide- the translation of ORF3c by Le et al. (78). The pro-
ally should have an unfavorable sequence context, posed coronavirus IRES elements are very different
allowing a proportion of ribosomes to bypass the first from those identified for picornaviruses; this is not
initiation codon and have the opportunity to initiate surprising, as the picornavirus IRES elements are for
protein synthesis at a second or third AUG codon with cap-independent translation, whereas coronaviruses
a more favorable sequence context (76). Most of the must function in concert with cap-dependent transla-
coronavirus and arterivirus sg mRNAs that are capable tion. If proteins are produced from coronavirus IRES-
of expressing two gene products are believed to do so like elements, there remains the possibility that these
as a result of leaky scanning. The sequence context of may also be expressed from other virus RNAs, gRNA
the initiation codons may control the amounts of the and larger sg mRNAs, which also contain these IRES-
proteins produced. In some cases the second ORF may like elements and downstream ORFs.
play an important role in the replication cycle of the Most of the coronavirus products, apart from
virus; e.g., MHV ORF5b yields the E protein, a struc- the E proteins of MHV and IBV, expressed from the
tural protein required for budding of virus particles. functionally bicistronic or tricistronic sg mRNAs repre-
The I ORF within the N protein gene of BCoV (121) sent small nonstructural genes of unknown function
or MHV (45) is also thought to be produced as the (for reviews, see references 19, 77, and 87). As out-
result of leaky scanning. Thus, expression of the MHV lined under “Structural and Accessory Protein Gene
E and I proteins as the second products of their Regions,” recent work has confirmed that these are
sg mRNAs indicates that bicistronic sg mRNAs con- accessory proteins and are not required for replica-
tribute to controlling the expression of these genes. tion per se. The observation that the synthesis of these
Another translation strategy for the control of products involves a variety of expression “tricks”
expression of virus gene products adopted by some raises the possibility that they play an important role
viruses, like picornaviruses, is the initiation of the in the viral replication cycle.
translation of an internal ORF using an internal ribo-
some entry site (IRES). An IRES element is usually
Copyright © 2007. ASM Press. All rights reserved.

REFERENCES
part of a complex RNA structure for cap-indepen-
dent translation. Its use allows viruses the opportu- 1. Almazan, F., C. Galan, and L. Enjuanes. 2004. The nucleopro-
nity to interfere with cap-dependent translation, tein is required for efficient coronavirus genome replication.
effectively inhibiting cellular protein synthesis with- J. Virol. 78:12683–12688.
2. Alonso, S., A. Izeta, I. Sola, and L. Enjuanes. 2002.
out affecting translation of virus-derived mRNAs.
Transcription regulatory sequences and mRNA expression lev-
However, as indicated above, coronaviruses, like all els in the coronavirus transmissible gastroenteritis virus.
members of the nidovirus order, employ the normal J. Virol. 76:1293–1308.
eukaryotic cap-dependent strategy for the translation 3. Archambault, D., A. Kheyar, A. A. de Vries, and P. J. Rottier.
of most virus-derived RNAs. 2006. The intraleader AUG nucleotide sequence context is
important for equine arteritis virus replication. Virus Genes
sg mRNA3 of the avian coronavirus IBV is a trcis-
33:59–68.
tronic mRNA, for the 3a, 3b, and 3c proteins; 3c is 4. Banerjee, S., K. Narayanan, T. Mizutani, and S. Makino.
the E protein. Studies have shown that 3a and 3b are 2002. Murine coronavirus replication-induced p38 mitogen-
in fact accessory proteins and not required for growth activated protein kinase activation promotes interleukin-6

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
42 BRITTON AND CAVANAGH

production and virus replication in cultured cells. J. Virol. 23. Cavanagh, D., K. Mawditt, M. Sharma, S. E. Drury, H. L.
76:5937–5948. Ainsworth, P. Britton, and R. E. Gough. 2001. Detection of a
5. Beerens, N., and E. J. Snijder. 2006. RNA signals in the 3 coronavirus from turkey poults in Europe genetically related
terminus of the genome of equine arteritis virus are required to infectious bronchitis virus of chickens. Avian Pathol.
for viral RNA synthesis. J. Gen. Virol. 87:1977–1983. 30:355–368.
6. Bhardwaj, K., L. Guarino, and C. C. Kao. 2004. The severe 24. Cavanagh, D., K. Mawditt, D. D. B. Welchman, P. Britton,
acute respiratory syndrome coronavirus nsp15 protein is an and R. E. Gough. 2002. Coronaviruses from pheasants
endoribonuclease that prefers manganese as a cofactor. (Phasianus colchicus) are genetically closely related to corona-
J. Virol. 78:12218–12224. viruses of domestic fowl (infectious bronchitis virus) and tur-
7. Boursnell, M. E. G., T. D. K. Brown, I. J. Foulds, P. F. Green, keys. Avian Pathol. 31:81–93.
F. M. Tomley, and M. M. Binns. 1987. Completion of the 25. Chen, H., T. Wurm, P. Britton, G. Brooks, and J. A. Hiscox.
sequence of the genome of the coronavirus avian infectious 2002. Interaction of the coronavirus nucleoprotein with nucle-
bronchitis virus. J. Gen. Virol. 68:57–77. olar antigens and the host cell. J. Virol. 76:5233–5250.
8. Bredenbeek, P. J., C. J. Pachuk, A. F. H. Noten, J. Charite, W. 26. Choi, J., Z. Xu, and J.-H. Ou. 2003. Triple decoding of hepa-
Luytjes, S. R. Weiss, and W. J. M. Spaan. 1990. The primary titis C virus RNA by programmed translational frameshifting.
structure and expression of the second open reading frame of Mol. Cell. Biol. 23:1489–1497.
the polymerase gene of the coronavirus MHV-A59—a highly 27. Cowley, J. A., C. M. Dimmock, K. M. Spann, and P. J. Walker.
conserved polymerase is expressed by an efficient ribosomal 2000. Gill-associated virus of Penaeus monodon prawns: an
frameshifting mechanism. Nucleic Acids Res. 18:1825–1832. invertebrate virus with ORF1a and ORF1b genes related to
9. Brian, D. A., and R. S. Baric. 2005. Coronavirus genome arteri- and coronaviruses. J. Gen. Virol. 81:1473–1484.
structure and replication. Curr. Top. Microbiol. Immunol. 28. Cowley, J. A., C. M. Dimmock, K. M. Spann, and P. J. Walker.
287:1–30. 2001. Gill-associated virus of Penaeus monodon prawns.
10. Brierley, I. 1995. Ribosomal frameshifting on viral RNAs. Molecular evidence for the first invertebrate nidovirus. Adv.
J. Gen. Virol. 76:1885–1892. Exp. Med. Biol. 494:43–48.
11. Brierley, I., M. E. G. Boursnell, M. M. Binns, B. Bilimoria, 29. Cowley, J. A., C. M. Dimmock, and P. J. Walker. 2002. Gill-
V. C. Blok, T. D. K. Brown, and S. C. Inglis. 1987. An efficient associated nidovirus of Penaeus monodon prawns transcribes
ribosomal frame-shifting signal in the polymerase encoding 3-coterminal subgenomic mRNAs that do not possess 5-
region of the coronavirus IBV. EMBO J. 6:3779–3785. leader sequences. J. Gen. Virol. 83:927–935.
12. Brierley, I., P. Digard, and S. C. Inglis. 1989. Characterization 30. Curtis, K. M., B. Yount, and R. S. Baric. 2002. Heterologous
of an efficient coronavirus ribosomal frameshifting signal— gene expression from transmissible gastroenteritis virus repli-
requirement for an RNA pseudoknot. Cell 57:537–547. con particles. J. Virol. 76:1422–1434.
13. Brierley, I., and F. J. Dos Ramos. 2006. Programmed ribosom- 31. Dalton, K., R. Casais, K. Shaw, K. Stirrups, S. Evans, P. Britton,
al frameshifting in HIV-1 and the SARS-CoV. Virus Res. T. D. Brown, and D. Cavanagh. 2001. cis-Acting sequences
119:29–42. required for coronavirus infectious bronchitis virus defective-
14. Brierley, I., A. J. Jenner, and S. C. Inglis. 1992. Mutational RNA replication and packaging. J. Virol. 75:125–133.
analysis of the “slippery-sequence” component of a coronavi- 32. De Groot, R. J., A. C. Andeweg, M. C. Horzinek, and W. J. M.
rus ribosomal frameshifting signal. J. Mol. Biol. 227:463– Spaan. 1988. Sequence analysis of the 3 end of the feline
479. coronavirus FIPV 79-1146 genome: comparison with the
15. Brierley, I., and S. Pennell. 2001. Structure and function of the genome of porcine coronavirus TGEV reveals large insertions.
stimulatory RNAs involved in programmed eukaryotic 1 Virology 167:370–376.
ribosomal frameshifting. Cold Spring Harbor Symp. Quant. 33. de Haan, C. A., P. S. Masters, X. Shen, S. Weiss, and P. J.
Biol. 66:233–248. Rottier. 2002. The group-specific murine coronavirus genes
16. Brierley, I., N. J. Rolley, A. J. Jenner, and S. C. Inglis. 1991. are not essential, but their deletion, by reverse genetics, is
Mutational analysis of the RNA pseudoknot component of a attenuating in the natural host. Virology 296:177–189.
coronavirus ribosomal frameshifting signal. J. Mol. Biol. 34. de Haan, C. A., H. Volders, C. A. Koetzner, P. S. Masters, and
220:889–902. P. J. Rottier. 2002. Coronaviruses maintain viability despite
17. Britton, P., and D. Cavanagh. 2007. Avian coronavirus dis- dramatic rearrangements of the strictly conserved genome
eases and infectious bronchitis vaccine development. In V. organization. J. Virol. 76:12491–12502.
Thiel (ed.), Coronaviruses: Molecular and Cellular Biology and 35. den Boon, J. A., M. F. Kleijnen, W. J. Spaan, and E. J. Snijder.
Diseases. Caister Academic Press, Norwich, United Kingdom. 1996. Equine arteritis virus subgenomic mRNA synthesis:
Copyright © 2007. ASM Press. All rights reserved.

18. Britton, P., C. L. Otin, J. M. M. Alonso, and F. Para. 1989. analysis of leader-body junctions and replicative-form RNAs.
Sequence of the coding regions from the 3.0Kb and 3.9Kb J. Virol. 70:4291–4298.
mRNA subgenomic species from a virulent isolate of transmis- 36. den Boon, J. A., E. J. Snijder, E. D. Chirnside, A. A. de Vries,
sible gastroenteritis virus. Arch. Virol. 105:165–178. M. C. Horzinek, and W. J. Spaan. 1991. Equine arteritis virus
19. Brown, T. D. K., and I. Brierly. 1995. The coronavirus non- is not a togavirus but belongs to the coronavirus like super-
structural proteins, p. 191–217. In S. G. Siddell (ed.), The family. J. Virol. 65:2910–2920.
Coronaviridae. Plenum Press, New York, NY. 37. de Vries, A. A. F., M. C. Horzinek, P. J. M. Rottier, and R. J.
20. Budzilowicz, C. J., and S. R. Weiss. 1987. In vitro synthesis of de Groot. 1997. The genome organisation of the Nidovirales:
two polypeptides from a nonstructural gene of coronavirus similarities and differences between Arteri-, Toro- and
mouse hepatitis virus strain A59. Virology 157:509–515. Coronaviruses. Semin. Virol. 8:33–47.
21. Casais, R., M. Davies, D. Cavanagh, and P. Britton. 2005. 38. Draker, R., R. L. Roper, M. Petric, and R. Tellier. 2006. The
Gene 5 of the avian coronavirus infectious bronchitis virus is complete sequence of the bovine torovirus genome. Virus Res.
not essential for replication. J. Virol. 79:8065–8078. 115:56–68.
22. Casais, R., V. Thiel, S. G. Siddell, D. Cavanagh, and P. Britton. 39. Edgil, D., and E. Harris. 2006. End-to-end communication in
2001. Reverse genetics system for the avian coronavirus infec- the modulation of translation by mammalian RNA viruses.
tious bronchitis virus. J. Virol. 75:12359–12369. Virus Res. 119:43–51.

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 43

40. Egli, M., S. Sarkhel, G. Minasov, and A. Rich. 2003. Structure the 3 untranslated region of the murine coronavirus genome.
and function of the ribosomal frameshifting pseudoknot RNA J. Virol. 74:6911–6921.
from beet western yellow virus. Helv. Chim. Acta 86:1709– 59. Hsue, B., and P. S. Masters. 1997. A bulged stem-loop struc-
1727. ture in the 3 untranslated region of the genome of the corona-
41. Eleouet, J. F., D. Rasschaert, P. Lambert, L. Levy, P. Vende, and virus mouse hepatitis virus is essential for replication.
H. Laude. 1995. Complete sequence (20 kilobases) of the J. Virol. 71:7567–7578.
polyprotein-encoding gene 1 of transmissible gastroenteritis 60. Ivanov, K. A., T. Hertzig, M. Rozanov, S. Bayer, V. Thiel,
virus. Virology 206:817–822. A. E. Gorbalenya, and J. Ziebuhr. 2004. Major genetic marker
42. Enjuanes, L. (ed.). 2005. Current Topics in Microbiology and of nidoviruses encodes a replicative endoribonuclease. Proc.
Immunology, vol. 287. Coronavirus Replication and Reverse Natl. Acad. Sci. USA 101:12694–12699.
Genetics. Springer, New York, NY. 61. Jacks, T., H. D. Madhani, F. R. Masiarz, and H. E. Varmus.
43. Enjuanes, L., F. Almazan, I. Sola, and S. Zuniga. 2006. 1988. Signals for ribosomal frameshifting in the Rous sarcoma
Biochemical aspects of coronavirus replication and virus-host virus gag-pol region. Cell 55:447–458.
interaction. Annu. Rev. Microbiol. 60:211–230. 62. Jacks, T., and H. E. Varmus. 1985. Expression of the Rous
44. Farabaugh, P. J. 1996. Programmed translational frameshift- sarcoma virus pol gene by ribosomal frameshifting. Science
ing. Microbiol. Rev. 60:103–134. 230:1237–1242.
45. Fischer, F., D. Peng, S. T. Hingley, S. R. Weiss, and P. S. 63. Jiang, B., S. S. Monroe, E. V. Koonin, S. E. Stine, and R. I.
Masters. 1997. The internal open reading frame within the Glass. 1993. RNA sequence of astrovirus: distinctive genomic
nucleocapsid gene of mouse hepatitis virus encodes a struc- organization and a putative retrovirus-like ribosomal frame-
tural protein that is not essential for viral replication. J. Virol. shifting signal that directs the viral replicase synthesis. Proc.
71:996–1003. Natl. Acad. Sci. USA 90:10539–10543.
46. Gingras, A. C., B. Raught, and N. Sonenberg. 1999. eIF4 ini- 64. Jonassen, C. M., T. O. Jonassen, and B. Grinde. 1998. A com-
tiation factors: effectors of mRNA recruitment to ribosomes mon RNA motif in the 3 end of the genomes of astroviruses,
and regulators of translation. Annu. Rev. Biochem. 68:913– avian infectious bronchitis virus and an equine rhinovirus.
963. J. Gen. Virol. 79:715–718.
47. Goebel, S. J., B. Hsue, T. F. Dombrowski, and P. S. Masters. 65. Jonassen, C. M., T. Kofstad, I. L. Larsen, A. Lovland, K.
2004. Characterization of the RNA components of a putative Handeland, A. Follestad, and A. Lillehaug. 2005. Molecular
molecular switch in the 3 untranslated region of the murine identification and characterization of novel coronaviruses
coronavirus genome. J. Virol. 78:669–682. infecting graylag geese (Anser anser), feral pigeons (Columbia
48. Goebel, S. J., J. Taylor, and P. S. Masters. 2004. The 3 cis-act- livia) and mallards (Anas platyrhynchos). J. Gen. Virol. 86:
ing genomic replication element of the severe acute respiratory 1597–1607.
syndrome coronavirus can function in the murine coronavirus 66. Kheyar, A., G. St-Laurent, and D. Archambault. 1996.
genome. J. Virol. 78:7846–7851. Sequence determination of the extreme 5 end of equine arte-
49. Gorbalenya, A. E., L. Enjuanes, J. Ziebuhr, and E. J. Snijder. ritis virus leader region. Virus Genes 12:291–295.
2006. Nidovirales: evolving the largest RNA virus genome. 67. Kienzle, T. E., S. Abraham, B. G. Hogue, and D. A. Brian.
Virus Res. 117:17–37. 1990. Structure and orientation of expressed bovine coronavi-
50. Gorbalenya, A. E., and E. V. Koonin. 1989. Viral proteins con- rus hemagglutinin-esterase protein. J. Virol. 64:1834–1838.
taining the purine NTP-binding sequence pattern. Nucleic 68. Kim, K. H., and S. A. Lommel. 1994. Identification and analy-
Acids Res. 17:8413–8440. sis of the site of 1 ribosomal frameshifting in red clover
51. Haijema, B. J., H. Volders, and P. J. Rottier. 2004. Live, atten- necrotic mosaic virus. Virology 200:574–582.
uated coronavirus vaccines through the directed deletion of 69. Kim, Y.-N., Y. S. Jeong, and S. Makino. 1993. Analysis of cis-
group-specific genes provide protection against feline infec- acting sequences essential for coronavirus defective interfering
tious peritonitis. J. Virol. 78:3863–3871. RNA replication. Virology 197:53–63.
52. Harger, J. W., A. Meskauskas, and J. D. Dinman. 2002. An 70. Kontos, H., S. Napthine, and I. Brierley. 2001. Ribosomal
‘integrated model’ of programmed ribosomal frameshifting. pausing at a frameshifter RNA pseudoknot is sensitive to read-
Trends Biochem. Sci. 27:448–454. ing phase but shows little correlation with frameshift efficien-
53. Herold, J., T. Raabe, B. Schelle-Prinz, and S. G. Siddell. 1993. cy. Mol. Cell. Biol. 21:8657–8670.
Nucleotide sequence of the human coronavirus 229E RNA 71. Kozak, M. 1981. Mechanism of mRNA recognition by eukary-
polymerase locus. Virology 195:680–691. otic ribosomes during initiation of protein synthesis. Curr.
54. Herold, J., and S. G. Siddell. 1993. An ‘elaborated’ pseu- Top. Microbiol. Immunol. 93:81–123.
Copyright © 2007. ASM Press. All rights reserved.

doknot is required for high frequency frameshifting during 72. Kozak, M. 1981. Possible role of flanking nucleotides in rec-
translation of HCV 229E polymerase mRNA. Nucleic Acids ognition of the AUG initiator codon by eukaryotic ribosomes.
Res. 21:5838–5842. Nucleic Acids Res. 9:5233–5252.
55. Hiscox, J. A., K. L. Mawditt, D. Cavanagh, and P. Britton. 73. Kozak, M. 1986. Regulation of protein synthesis in virus-
1995. Investigation of the control of coronavirus subgenomic infected animal cells. Adv. Virus Res. 31:229–292.
mRNA transcription by using T7-generated negative-sense 74. Kozak, M. 1986. Bifunctional messenger RNAs in eukaryotes.
RNA transcripts. J. Virol. 69:6219–6227. Cell 47:481–483.
56. Hiscox, J. A., T. Wurm, L. Wilson, P. Britton, D. Cavanagh, and 75. Kozak, M. 1987. At least six nucleotides preceding the AUG
G. Brooks. 2001. The coronavirus infectious bronchitis virus initiator codon enhance translation in mammalian cells.
nucleoprotein localizes to the nucleolus. J. Virol. 75:506–512. J. Mol. Biol. 196:947–950.
57. Hodgson, T., P. Britton, and D. Cavanagh. 2006. Neither the 76. Kozak, M. 1989. The scanning model for translation: an
RNA nor the proteins of open reading frames 3a and 3b of the update. J. Cell Biol. 108:229–241.
coronavirus infectious bronchitis virus are essential for repli- 77. Lai, M. M., and D. Cavanagh. 1997. The molecular biology of
cation. J. Virol. 80:296–305. coronaviruses. Adv. Virus Res. 48:1–100.
58. Hsue, B., T. Hartshorne, and P. S. Masters. 2000. 78. Le, S. Y., N. Sonenberg, and J. V. Maizel, Jr. 1994. Distinct
Characterization of an essential RNA secondary structure in structural elements and internal entry of ribosomes in mRNA3

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
44 BRITTON AND CAVANAGH

encoded by infectious bronchitis virus. Virology 198:405– 94. Molenkamp, R., S. Greve, W. J. Spaan, and E. J. Snijder.
411. 2000. Efficient homologous RNA recombination and require-
79. Lee, H.-J., C.-K. Shieh, A. E. Gorbalenya, E. V. Koonin, N. La ment for an open reading frame during replication of equine
Monica, J. Tuler, A. Bagdzhadzhyan, and M. M. C. Lai. 1991. arteritis virus defective interfering RNAs. J. Virol. 74:9062–
The complete sequence (22 kilobases) of murine coronavirus 9070.
gene 1 encoding the putative proteases and RNA polymerase. 95. Molenkamp, R., H. van Tol, B. C. Rozier, Y. van Der Meer,
Virology 180:567–582. W. J. Spaan, and E. J. Snijder. 2000. The arterivirus replicase
80. Leparc-Goffart, I., S. T. Hingley, M. M. Chua, X. Jiang, E. Lavi, is the only viral protein required for genome replication and
and S. R. Weiss. 1997. Altered pathogenesis of a mutant of the subgenomic mRNA transcription. J. Gen. Virol. 81:2491–
murine coronavirus MHV-A59 is associated with a Q159L amino 2496.
acid substitution in the spike protein. Virology 239:1–10. 96. Namy, O., S. J. Moran, D. I. Stuart, R. J. Gilbert, and I.
81. Lin, Y. J., and M. M. C. Lai. 1993. Deletion mapping of a Brierley. 2006. A mechanical explanation of RNA pseu-
mouse hepatitis virus defective interfering RNA reveals the doknot function in programmed ribosomal frameshifting.
requirement of an internal and discontiguous sequence for Nature 441:244–247.
replication. J. Virol. 67:6110–6118. 97. Ortego, J., I. Sola, F. Almazan, J. E. Ceriani, C. Riquelme, M.
82. Lin, Y. J., C. L. Liao, and M. M. Lai. 1994. Identification of Balasch, J. Plana, and L. Enjuanes. 2003. Transmissible gas-
the cis-acting signal for minus-strand RNA synthesis of a troenteritis coronavirus gene 7 is not essential but influences
murine coronavirus: implications for the role of minus-strand in vivo virus replication and virulence. Virology 308:13–22.
RNA in RNA replication and transcription. J. Virol. 68:8131– 98. Pasternak, A. O., A. P. Gultyaev, W. J. Spaan, and E. J.
8140. Snijder. 2000. Genetic manipulation of arterivirus alternative
83. Liu, D. X., D. Cavanagh, P. Green, and S. C. Inglis. 1991. A mRNA leader-body junction sites reveals tight regulation of
polycistronic mRNA specified by the coronavirus infectious structural protein expression. J. Virol. 74:11642–11653.
bronchitis virus. Virology 184:531–544. 99. Pasternak, A. O., W. J. Spaan, and E. J. Snijder. 2004.
84. Liu, D. X., and S. C. Inglis. 1992. Identification of two new Regulation of relative abundance of arterivirus subgenomic
polypeptides encoded by mRNA5 of the coronavirus infec- mRNAs. J. Virol. 78:8102–8113.
tious bronchitis virus. Virology 186:342–347. 100. Pasternak, A. O., E. van den Born, W. J. M. Spaan, and E. J.
85. Liu, D. X., and S. C. Inglis. 1992. Internal entry of ribosomes Snijder. 2003. The stability of the duplex between sense and
on a tricistronic mRNA encoded by infectious bronchitis virus. antisense transcription-regulating sequences is a crucial fac-
J. Virol. 66:6143–6154. tor in arterivirus subgenomic mRNA synthesis. J. Virol.
86. Lopinski, J. D., J. D. Dinman, and J. A. Bruenn. 2000. Kinetics 77:1175–1183.
of ribosomal pausing during programmed 1 translational 101. Plant, E. P., and J. D. Dinman. 2006. Comparative study of
frameshifting. Mol. Cell. Biol. 20:1095–1103. the effects of heptameric slippery site composition on 1
87. Luytjes, W. 1995. Coronavirus gene expression: genome frameshifting among different eukaryotic systems. RNA
organisation and protein synthesis, p. 33–54. In S. G. Siddell 12:666–673.
(ed.), The Coronaviridae. Plenum Press, New York, NY. 102. Plant, E. P., K. L. M. Jacobs, J. W. Harger, A. Meskauskas,
88. Maines, T. R., M. Young, N. N. Dinh, and M. A. Brinton. J. L. Jacobs, J. L. Baxter, A. N. Petrov, and J. D. Dinman. 2003.
2005. Two cellular proteins that interact with a stem loop in The 9-Å solution: how mRNA pseudoknots promote efficient
the simian hemorrhagic fever virus 3 ()NCR RNA. Virus programmed 1 ribosomal frameshifting. RNA 9:168–174.
Res. 109:109–124. 103. Pleij, C. W. A., and L. Bosch. 1989. RNA pseudoknots:
89. Marra, M. A., S. J. Jones, C. R. Astell, R. A. Holt, A. Brooks- structure, detection and prediction. Methods Enzymol.
Wilson, Y. S. Butterfield, J. Khattra, J. K. Asano, S. A. Barber, 180:289–303.
S. Y. Chan, A. Cloutier, S. M. Coughlin, D. Freeman, N. Girn, 104. Posthuma, C. C., D. D. Nedialkova, J. C. Zevenhoven-
O. L. Griffith, S. R. Leach, M. Mayo, H. McDonald, S. B. Dobbe, J. H. Blokhuis, A. E. Gorbalenya, and E. J. Snijder.
Montgomery, P. K. Pandoh, A. S. Petrescu, A. G. Robertson, 2006. Site-directed mutagenesis of the nidovirus replicative
J. E. Schein, A. Siddiqui, D. E. Smailus, J. M. Stott, G. S. Yang, endoribonuclease NendoU exerts pleiotropic effects on the
F. Plummer, A. Andonov, H. Artsob, N. Bastien, K. Bernard, arterivirus life cycle. J. Virol. 80:1653–1661.
T. F. Booth, D. Bowness, M. Czub, M. Drebot, L. Fernando, 105. Prufer, D., E. Tacke, J. Schmitz, B. Kull, A. Kaufmann, and
R. Flick, M. Garbutt, M. Gray, A. Grolla, S. Jones, H. W. Rohde. 1992. Ribosomal frameshifting in plants: a novel
Feldmann, A. Meyers, A. Kabani, Y. Li, S. Normand, U. signal directs the 1 frameshift in the synthesis of the puta-
Stroher, G. A. Tipples, S. Tyler, R. Vogrig, D. Ward, B. Watson, tive viral replicase of potato leafroll luteovirus. EMBO J. 11:
Copyright © 2007. ASM Press. All rights reserved.

R. C. Brunham, M. Krajden, M. Petric, D. M. Skowronski, C. 1111–1117.


Upton, and R. L. Roper. 2003. The genome sequence of the 106. Putics, A., W. Filipowicz, J. Hall, A. E. Gorbalenya, and J.
SARS-associated coronavirus. Science 300:1399–1404. Ziebuhr. 2005. ADP-ribose-1-monophosphatase: a con-
90. Masters, P. S. 2006. The molecular biology of coronaviruses. served coronavirus enzyme that is dispensable for viral repli-
Adv. Virus Res. 66:193–292. cation in tissue culture. J. Virol. 79:12721–12731.
91. Meier, C., A. R. Aricescu, R. Assenberg, R. T. Aplin, R. J. C. 107. Putics, A., A. E. Gorbalenya, and J. Ziebuhr. 2006.
Gilbert, J. M. Grimes, and D. I. Stuart. 2006. The crystal Identification of protease and ADP-ribose 1-monophospha-
structure of ORF-9b, a lipid binding protein from the SARS tase activities associated with transmissible gastroenteritis
coronavirus. Structure 14:1157–1165. virus non-structural protein 3. J. Gen. Virol. 87:651–656.
92. Mizutani, T., S. Fukushi, M. Saijo, I. Kurane, and S. Morikawa. 108. Raabe, T., and S. Siddell. 1989. Nucleotide sequence of the
2004. Phosphorylation of p38 MAPK and its downstream tar- human coronavirus HCV 229E mRNA 4 and mRNA 5
gets in SARS coronavirus-infected cells. Biochem. Biophys. unique regions. Nucleic Acids Res. 17:6387.
Res. Commun. 319:1228–1234. 109. Raman, S., P. Bouma, G. D. Williams, and D. A. Brian. 2003.
93. Mohr, I. 2006. Phosphorylation and dephosphorylation Stem-loop III in the 5 untranslated region is a cis-acting ele-
events that regulate viral mRNA translation. Virus Res. ment in bovine coronavirus defective interfering RNA repli-
119:89–99. cation. J. Virol. 77:6720–6730.

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
CHAPTER 3 • GENOME ORGANIZATION AND EXPRESSION MECHANISMS 45

110. Raman, S., and D. A. Brian. 2005. Stem-loop IV in the 5 128. Smits, S. L., E. J. Snijder, and R. J. de Groot. 2006.
untranslated region is a cis-acting element in bovine corona- Characterization of a torovirus main proteinase. J. Virol.
virus defective interfering RNA replication. J. Virol. 79: 80:4157–4167.
12434–12446. 129. Snijder, E. J., P. J. Bredenbeek, J. C. Dobbe, V. Thiel, J.
111. Reed, M. L., B. K. Dove, R. M. Jackson, R. Collins, G. Ziebuhr, L. L. Poon, Y. Guan, M. Rozanov, W. J. Spaan, and
Brooks, and J. A. Hiscox. 2006. Delineation and modelling A. E. Gorbalenya. 2003. Unique and conserved features of
of a nucleolar retention signal in the coronavirus nucleocap- genome and proteome of SARS-coronavirus, an early split-
sid protein. Traffic 7:1–16. off from the coronavirus group 2 lineage. J. Mol. Biol.
112. Robertson, M. P., H. Igel, R. Baertsch, D. Haussler, M. Ares, 331:991–1004.
and W. G. Scott. 2005. The structure of a rigorously con- 130. Snijder, E. J., J. A. den Boon, P. J. Bredenbeek, M. C.
served RNA element within the SARS virus genome. PLoS Horzinek, R. Rijnbrand, and W. Spaan. 1990. The carboxyl-
Biol. 3:e5. terminal part of the putative Berne virus polymerase is
113. Rowland, R. R., R. Kerwin, C. Kuckleburg, A. Sperlich, expressed by ribosomal frameshifting and contains sequence
and D. A. Benfield. 1999. The localisation of porcine repro- motifs which indicate that toro- and coronaviruses are evolu-
ductive and respiratory syndrome virus nucleocapsid protein tionarily related. Nucleic Acids Res. 18:4535–4542.
to the nucleolus of infected cells and identification of a poten- 131. Snijder, E. J., H. van Tol, K. W. Pedersen, M. J. Raamsman,
tial nucleolar localisation signal sequence. Virus Res. 64: and A. A. de Vries. 1999. Identification of a novel structural
1–12. protein of arteriviruses. J. Virol. 73:6335–6345.
114. Saikatendu, K. S., J. S. Joseph, V. Subramanian, T. Clayton, 132. Sola, I., S. Alonso, S. Zuniga, M. Balasch, J. Plana-Duran,
M. Griffith, K. Moy, J. Velasquez, B. W. Neuman, M. J. and L. Enjuanes. 2003. Engineering the transmissible gastro-
Buchmeier, R. C. Stevens, and P. Kuhn. 2005. Structural enteritis virus genome as an expression vector inducing lacto-
basis of severe acute respiratory syndrome coronavirus genic immunity. J. Virol. 77:4357–4369.
ADP-ribose-1-phosphate dephosphorylation by a conserved 133. Sola, I., J. L. Moreno, S. Zuniga, S. Alonso, and L. Enjuanes.
domain of nsP3. Structure 13:1665–1675. 2005. Role of nucleotides immediately flanking the
115. Sawicki, D., T. Wang, and S. Sawicki. 2001. The RNA struc- transcription-regulating sequence core in coronavirus subge-
tures engaged in replication and transcription of the A59 nomic mRNA synthesis. J. Virol. 79:2506–2516.
strain of mouse hepatitis virus. J. Gen. Virol. 82:385–396. 134. Somogyi, P., A. J. Jenner, I. Brierley, and S. C. Inglis. 1993.
116. Sawicki, S. G., and D. L. Sawicki. 1995. Coronaviruses use Ribosomal pausing during translation of an RNA pseu-
discontinuous extension for synthesis of subgenome-length doknot. Mol. Cell. Biol. 13:6931–6940.
negative strands. Adv. Exp. Med. Biol. 380:499–506. 135. Spagnolo, J. F., and B. G. Hogue. 2000. Host protein interac-
117. Sawicki, S. G., and D. L. Sawicki. 1998. A new model for tions with the 3 end of bovine coronavirus RNA and the
coronavirus transcription. Adv. Exp. Med. Biol. 440:215–219. requirement of the poly(A) tail for coronavirus defective
118. Sawicki, S. G., and D. L. Sawicki. 2005. Coronavirus tran- genome replication. J. Virol. 74:5053–5065.
scription: a perspective. Curr. Top. Microbiol. Immunol. 287: 136. Stirrups, K., K. Shaw, S. Evans, K. Dalton, D. Cavanagh,
31–55. and P. Britton. 2000. Leader switching occurs during the
119. Sawicki, S. G., D. L. Sawicki, and S. G. Siddell. 2007. A con- rescue of defective RNAs by heterologous strains of the
temporary view of coronavirus transcription. J. Virol. 81: coronavirus infectious bronchitis virus. J. Gen. Virol. 81:
20–29. 791–801.
120. Schelle, B., N. Karl, B. Ludewig, S. G. Siddell, and V. Thiel. 137. Sturman, L. S., and K. V. Holmes. 1983. The molecular biol-
2005. Selective replication of coronavirus genomes that ogy of coronaviruses. Adv. Virus Res. 28:35–112.
express nucleocapsid protein. J. Virol. 79:6620–6630. 138. Tahara, S. M., T. A. Dietlin, C. C. Bergmann, G. W. Nelson,
121. Senanayake, S. D., and D. A. Brian. 1997. Bovine coronavi- S. Kyuwa, R. P. Anthony, and S. A. Stohlman. 1994.
rus I protein synthesis follows ribosomal scanning on the Coronavirus translational regulation: leader affects mRNA
bicistronic N mRNA. Virus Res. 48:101–105. efficiency. Virology 202:621–630.
122. Senanayake, S. D., and D. A. Brian. 1999. Translation from 139. Tahara, S. M., T. A. Dietlin, G. W. Nelson, S. A. Stohlman,
the 5 untranslated region (UTR) of mRNA 1 is repressed, and D. J. Manno. 1998. Mouse hepatitis virus nucleocapsid
but that from the 5 UTR of mRNA 7 is stimulated in coro- protein as a translational effector of viral mRNAs. Adv. Exp.
navirus-infected cells. J. Virol. 73:8003–8009. Med. Biol. 440:313–318.
123. Senanayake, S. D., M. A. Hofmann, J. L. Maki, and D. A. 140. Thiel, V., K. A. Ivanov, A. Putics, T. Hertzig, B. Schelle, S.
Brian. 1992. The nucleocapsid protein gene of bovine coro- Bayer, B. Weissbrich, E. J. Snijder, H. Rabenau, H. W. Doerr,
Copyright © 2007. ASM Press. All rights reserved.

navirus is bicistronic. J. Virol. 66:5277–5283. A. E. Gorbalenya, and J. Ziebuhr. 2003. Mechanisms and
124. Seybert, A., C. C. Posthuma, L. C. van Dinten, E. J. Snijder, enzymes involved in SARS coronavirus genome expression.
A. E. Gorbalenya, and J. Ziebuhr. 2005. A complex zinc fin- J. Gen. Virol. 84:2305–2315.
ger controls the enzymatic activities of nidovirus helicases. 141. Thiel, V., and S. G. Siddell. 2005. Reverse genetics of corona-
J. Virol. 79:696–704. viruses using vaccinia virus vectors. Curr. Top. Microbiol.
125. Siddell, S., J. Ziebuhr, and E. J. Snijder. 2005. Coronaviruses, Immunol. 287:199–227.
toroviruses and arteriviruses, p. 823–856. In B. W. J. Mahy 142. Tijms, M. A., Y. van der Meer, and E. J. Snijder. 2002.
and V. ter Meulen (ed.), Topley and Wilson’s Microbiology Nuclear localization of nonstructural protein 1 and nucleo-
and Microbial Infections: Virology. Hodder Arnold, London, capsid protein of equine arteritis virus. J. Gen. Virol. 83:
United Kingdom. 795–800.
126. Siddell, S. G. (ed.). 1995. The Coronaviridae. Plenum, New 143. Tu, C., T. H. Tzeng, and J. A. Bruenn. 1992. Ribosomal
York, NY. frameshifting impeded at a pseudoknot required for frame-
127. Skinner, M. A., D. Ebner, and S. G. Siddell. 1985. Coronavirus shifting. Proc. Natl. Acad. Sci. USA 89:8636–8640.
MHV-JHM mRNA 5 has a sequence arrangement which 144. van Aken, D., J. Zevenhoven-Dobbe, A. E. Gorbalenya, and
potentially allows translation of a second, downstream open E. J. Snijder. 2006. Proteolytic maturation of replicase poly-
reading frame. J. Gen. Virol. 66:581–592. protein pp1a by the nsp4 main proteinase is essential for

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.
46 BRITTON AND CAVANAGH

equine arteritis virus replication and includes internal cleav- 152. Williams, G. D., R. Y. Chang, and D. A. Brian. 1999. A phy-
age of nsp7. J. Gen. Virol. 87:3473–3482. logenetically conserved hairpin-type 3 untranslated region
145. van den Born, E., A. P. Gultyaev, and E. J. Snijder. 2004. pseudoknot functions in coronavirus RNA replication.
Secondary structure and function of the 5-proximal region J. Virol. 73:8349–8355.
of the equine arteritis virus RNA genome. RNA 10:424–437. 153. Wurm, T., H. Chen, T. Hodgson, P. Britton, G. Brooks, and
146. van Dinten, L. C., H. van Tol, A. E. Gorbalenya, and E. J. J. A. Hiscox. 2001. Localization to the nucleolus is a com-
Snijder. 2000. The predicted metal-binding region of the mon feature of coronavirus nucleoproteins, and the protein
arterivirus helicase protein is involved in subgenomic mRNA may disrupt host cell division. J. Virol. 75:9345–9356.
synthesis, genome replication, and virion biogenesis. J. Virol. 154. Xu, Z., J. Choi, T. S. Yen, W. Lu, A. Strohecker, S.
74:5213–5223. Govindarajan, D. Chien, M. J. Selby, and J. Ou. 2001.
147. van Marle, G., L. C. van Dinten, W. J. Spaan, W. Luytjes, and Synthesis of a novel hepatitis C virus protein by ribosomal
E. J. Snijder. 1999. Characterization of an equine arteritis frameshift. EMBO J. 20:3840–3848.
virus replicase mutant defective in subgenomic mRNA syn- 155. Yamanaka, M., T. Crisp, R. Brown, and B. Dale.
thesis. J. Virol. 73:5274–5281. 1998. Nucleotide sequence of the inter-structural gene region
148. van Vliet, A. L. W., S. L. Smits, P. J. M. Rottier, and R. J. de of feline infectious peritonitis virus. Virus Genes 16:317–
Groot. 2002. Discontinuous and non-discontinuous subge- 318.
nomic RNA transcription in a nidovirus. EMBO J. 21:6571– 156. You, J., B. K. Dove, L. Enjuanes, M. L. DeDiego, E. Alvarez,
6580. G. Howell, P. Heinen, M. Zambon, and J. A. Hiscox. 2005.
149. Verheije, M. H., R. C. Olsthoorn, M. V. Kroese, P. J. Rottier, Subcellular localization of the severe acute respiratory syn-
and J. J. Meulenberg. 2002. Kissing interaction between 3 drome coronavirus nucleocapsid protein. J. Gen. Virol.
noncoding and coding sequences is essential for porcine 86:3303–3310.
arterivirus RNA replication. J. Virol. 76:1521–1526. 157. Youn, S., J. L. Leibowitz, and E. W. Collisson. 2005. In vitro
150. Wang, A. L., H. M. Yang, K. A. Shen, and C. C. Wang. 1993. assembled, recombinant infectious bronchitis viruses demon-
Giardiavirus double-stranded RNA genome encodes a capsid strate that the 5a open reading frame is not essential for rep-
polypeptide and a gag-pol-like fusion protein by a transla- lication. Virology 332:206–215.
tion frameshift. Proc. Natl. Acad. Sci. USA 90:8595–8599. 158. Zúñiga, S., I. Sola, S. Alonso, and L. Enjuanes. 2004.
151. Wang, Y., and X. Zhang. 2000. The leader RNA of coronavi- Sequence motifs involved in the regulation of discontinuous
rus mouse hepatitis virus contains an enhancer-like element for coronavirus subgenomic RNA synthesis. J. Virol. 78:980–
subgenomic mRNA transcription. J. Virol. 74:10571–10580. 994.
Copyright © 2007. ASM Press. All rights reserved.

Nidoviruses, edited by Stanley Perlman, et al., ASM Press, 2007. ProQuest Ebook Central, http://ebookcentral.proquest.com/lib/bibliotecausta-ebooks/detail.action?docID=476468.
Created from bibliotecausta-ebooks on 2020-04-21 10:53:30.

You might also like