ArticlePDF AvailableLiterature Review

Abstract and Figures

Although they induce symptoms in plants similar to those accompanying virus infections, viroids have unique structural, functional, and evolutionary characteristics. They are composed of a small, nonprotein-coding, single-stranded, circular RNA, with autonomous replication. Viroid species are clustered into the families Pospiviroidae and Avsunviroidae, whose members replicate (and accumulate) in the nucleus and chloroplast, respectively. Viroids replicate in three steps through an RNA-based rolling-circle mechanism: synthesis of longer-than-unit strands catalyzed by host RNA polymerases; processing to unit-length, which in the family Avsunviroidae is mediated by hammerhead ribozymes; and circularization. Within the initially infected cells, viroid RNA must move to its replication organelle, with the resulting progeny then invading adjacent cells through plasmodesmata and reaching distal parts via the vasculature. To carry out these movements, viroids must interact with host factors. The mature viroid RNA could be the primary pathogenic effector or, alternatively, viroids could exert their pathogenic effects via RNA silencing.
Structural features of viroids. (a) Rod-like secondary structure of members of the family Pospiviroidae. Domains C (central), P (pathogenic), V (variable), and T L and T R (terminal left and right, respectively) are indicated, as well as the CCR (central conserved region, here displayed for the genus Pospiviroid), TCR (terminal conserved region, present in the genera Pospi-and Apscaviroid, and in the two largest members of the genus Coleviroid) and TCH (terminal conserved hairpin, present in the genera Hostu-and Cocadviroid). Arrows indicate flanking sequences that together with the upper CCR strand form a hairpin, and the S-shaped line connects the residues linked after UV irradiation as a consequence of forming part of the loop E. (b) Quasi-rod-like and branched secondary structures of ASBVd and PLMVd, respectively (family Avsunviroidae). Sequences conserved in most natural hammerhead structures are shown within boxes with blue and white backgrounds for () and () polarities, respectively. Broken oval in PLMVd denotes a kissing-loop interaction. (Inset) PLMVd () hammerhead structure represented according to the original scheme (left) and to X-ray crystallography data obtained with artificial hammerhead structures (right), in which a proposed tertiary interaction between loops 1 and 2 enhancing the catalytic activity is indicated. Nucleotides conserved in most natural hammerhead structures are depicted as above. Arrows mark the self-cleavage site, and continuous and broken lines denote Watson-Crick and noncanonical pairs, respectively.
… 
Content may be subject to copyright.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
10.1146/annurev.phyto.43.040204.140243
Annu. Rev. Phytopathol. 2005. 43:117–39
doi: 10.1146/annurev.phyto.43.040204.140243
Copyright c
2005 by Annual Reviews. All rights reserved
First published online as a Review in Advance on February 7, 2005
VIROIDS AND VIROID-HOST INTERACTIONS
Ricardo Flores,1Carmen Hern´
andez,1
A. Emilio Mart´
ınez de Alba,1Jos ´
e-Antonio Dar `
os,1
and Francesco Di Serio2
1Instituto de Biolog´
ıa Molecular y Celular de Plantas (UPV-CSIC), Universidad
Polit´
ecnica de Valencia, Valencia 46022, Spain; email: rflores@ibmcp.upv.es,
cahernan@ibmcp.upv.es, aemarti@ibmcp.upv.es, jadaros@ibmcp.upv.es
2Istituto di Virologia Vegetale (CNR), Dipartimento di Protezione delle Piante e
Microbiologia Applicata, Universit`
adegli Studi di Bari, Bari 70126, Italy;
email: f.diserio@ba.ivv.cnr.it
KeyWords rolling-circle replication, catalytic RNAs, hammerhead ribozymes,
RNA silencing, cross-protection
Abstract Although they induce symptoms in plants similar to those accompa-
nying virus infections, viroids have unique structural, functional, and evolutionary
characteristics. They are composed of a small, nonprotein-coding, single-stranded, cir-
cular RNA, with autonomous replication. Viroid species are clustered into the families
Pospiviroidae and Avsunviroidae, whose members replicate (and accumulate) in the
nucleus and chloroplast, respectively. Viroids replicate in three steps through an RNA-
based rolling-circle mechanism: synthesis of longer-than-unit strands catalyzed by host
RNA polymerases; processing to unit-length, which in the family Avsunviroidae is me-
diated by hammerhead ribozymes; and circularization. Within the initially infected
cells, viroid RNA must move to its replication organelle, with the resulting progeny
then invading adjacent cells through plasmodesmata and reaching distal parts via the
vasculature. To carry out these movements, viroids must interact with host factors. The
mature viroid RNA could be the primary pathogenic effector or, alternatively, viroids
could exert their pathogenic effects via RNA silencing.
INTRODUCTION
The closest reference term to viroid is virus, reflecting the intimate historical links
between viruses, discovered at the end of the nineteenth century in plants, and vi-
roids, also discovered in plants some 70 years later. The first viroid, Potato spindle
tuber viroid (PSTVd) (25, 27), was identified during an attempt to characterize the
virus presumed to cause a potato disease. These and subsequent results obtained
for the same disease (97) as well as for another disease that was also presumed
to have a virus etiology but which turned out to be induced by Citrus exocor-
tis viroid (CEVd) (86, 96), established the viroid concept on solid experimental
ground. Viroids and viruses cannot, in principle, be discriminated according to the
0066-4286/05/0908-0117$20.00 117
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
118 FLORES ET AL.
phenotypic effects that they incite in their hosts, which range from severe symptoms
to latent infections. However, both entities differ radically in structure, function,
and evolutionary origin, and if they are collected together in taxonomic schemes
like the one endorsed by the International Committee for Taxonomy of Viruses
(ICTV) (36), it is for historical and practical reasons. (Prions are also included in
this scheme, even though there is an even more drastic difference between them
and viruses). In this review we focus on the interactions of viroids with their host
plants but first we briefly describe the properties of these unique pathogens and
their classification (for a detailed coverage of different aspects of viroids, see 42).
TAXONOMY OF THE VIROIDAE:
FAMILIES, GENERA, SPECIES
Figure 1 summarizes the structural characteristics of viroids. Because their mech-
anisms of replication and pathogenesis are discussed in detail below, here we
only highlight other properties that distinguish them from viruses. Viroids are
tiny single-stranded RNAs (246–401 nt), approximately tenfold smaller than the
genome of the smallest RNA viruses; they have a circular structure and high de-
gree of self-complementarity that promotes compact folding (87). As a functional
reflection of these singular features, viroids, in contrast to viruses, do not appear
to code for specific proteins (reviewed in 26), hence they must rely almost entirely
on host factors to complete their infectious cycle. This means that the parasitism
developed by viroids and viruses also differs: Viruses and viroids can essentially
be regarded as parasites of the translation and transcriptional apparatus of their
hosts, respectively. However, some viroids are catalytic RNAs and “code” for
hammerhead ribozymes that mediate the self-cleavage of the multimeric RNAs
generated in their replication through a rolling-circle mechanism. Apart from cer-
tain plant satellite RNAs and the RNA of human hepatitis delta virus (reviewed
in 35, 101), which resemble viroids in their circular structure and rolling-circle
replication mode mediated by ribozymes, no other virus-related RNAs have been
characterized as catalytic RNAs. This property constitutes the strongest argument
in support of the idea that viroids may have an ancient evolutionary origin inde-
pendent of viruses, going back to the RNA world postulated to have preceded the
present world on Earth based on DNA and proteins (reviewed in 26).
Most of the nearly 30 viroid species known (36) (Table 1) belong to the family
Pospiviroidae, type species PSTVd (25, 41), and adopt in vitro a rod-like or quasi-
rod-like secondary structure of minimal free energy (29, 87) with five structural-
functional domains (52, 89) (Figure 1a). The central conserved region (CCR),
within the C domain, is formed by two stretches of conserved nucleotides, in
which those of the upper strand are flanked by an inverted repeat. Depending of
the nature of the CCR, and on the presence or absence of a terminal conserved
region (TCR) and a terminal conserved hairpin (TCH), members of this family are
allocated to five genera (36) (Table 1). The other four viroids, Avocado sunblotch
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 119
TABLE 1 Viroid species with their abbreviations, accession numbers of typical sequence
variants, sizes, and genus and family to which they belong
Viroid species Abbreviation Accession Size (nt) Genus Family
Potato spindle tuber PSTVd V01465 359 Pospiviroid Pospiviroidae
Tomato chlorotic
dwarf
TCDVd AF162131 360 Pospiviroid Pospiviroidae
Mexican papita MPVd L78454 360 Pospiviroid Pospiviroidae
Tomato planta
macho
TPMVd K00817 360 Pospiviroid Pospiviroidae
Citrus exocortis CEVd M34917 371 Pospiviroid Pospiviroidae
Chrysanthemum
stunt
CSVd V01107 356 Pospiviroid Pospiviroidae
Tomato apical stunt TASVd K00818 360 Pospiviroid Pospiviroidae
Iresine 1 IrVd-1 X95734 370 Pospiviroid Pospiviroidae
Columnea latent CLVd X15663 370 Pospiviroid Pospiviroidae
Hop stunt HSVd X00009 297 Hostuviroid Pospiviroidae
Coconut
cadang-cadang
CCCVd J02049 246 Cocadviroid Pospiviroidae
Coconut tinangaja CTiVd M20731 254 Cocadviroid Pospiviroidae
Hop latent HLVd X07397 256 Cocadviroid Pospiviroidae
Citrus IV CVd-IV X14638 284 Cocadviroid Pospiviroidae
Apple scar skin ASSVd M36646 329 Apscaviroid Pospiviroidae
Citrus III CVd-III AF184147 294 Apscaviroid Pospiviroidae
Apple dimple fruit ADFVd X99487 306 Apscaviroid Pospiviroidae
Grapevine yellow
speckle 1
GVYSd-1 X06904 367 Apscaviroid Pospiviroidae
Grapevine yellow
speckle 2
GVYSd-2 J04348 363 Apscaviroid Pospiviroidae
Citrus bent leaf CBLVd M74065 318 Apscaviroid Pospiviroidae
Pear blister canker PBCVd D12823 315 Apscaviroid Pospiviroidae
Australian
grapevine
AGVd X17101 369 Apscaviroid Pospiviroidae
Coleus blumei 1 CbVd-1 X52960 248 Coleviroid Pospiviroidae
Coleus blumei 2 CbVd-2 X95365 301 Coleviroid Pospiviroidae
Coleus blumei 3 CbVd-3 X95364 361 Coleviroid Pospiviroidae
Avocado sunblotch ASBVd J02020 247 Avsunviroid Avsunviroidae
Peach latent mosaic PLMVd M83545 337 Pelamoviroid Avsunviroidae
Chrysanthemum
chlorotic mottle
CChMVd Y14700 399 Pelamoviroid Avsunviroidae
Eggplant latentELVd AJ536613 333 Elaviroid Avsunviroidae
Pending ICTV approval; whether Apple fruit crinkle and Citrus viroid original source should be considered as new viroid
species of genus Apscaviroid is also pending.
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
120 FLORES ET AL.
viroid (ASBVd) (47), Peach latent mosaic viroid (PLMVd) (45), Chrysanthemum
chlorotic mottle viroid (CChMVd) (65), and Eggplant latent viroid (ELVd) (30),
do not have the conserved CCR, TCR, and TCH motifs but, remarkably, both their
polarity strands self-cleave through hammerhead ribozymes; they form the second
family, Avsunviroidae (reviewed in 33), whose type species is ASBVd (formal
inclusion of ELVd in this family is pending ICTV approval) (Figure 1b). Apart
from the core nucleotides conserved in their hammerhead structures, no extensive
sequence similarities exist between them, but PLMVd and CChMVd are grouped
in one genus because of their branched secondary structure (21, 45, 65), which is
stabilized by a pseudoknot (10; S. Gago, M. De la Pe˜na & R. Flores, unpublished
results) (Figure 1b), and their insolubility in2MLiCl (65). ASBVd, the only
viroid with a high A +U content (62%) (47), forms a monospecific genus, and
ELVd, whose properties fall between those of the members of the other two genera,
has been proposed to constitute its own genus (30). This classification scheme is
further supported by phylogenetic reconstructions with entire viroid sequences
(36) and by the different subcellular replication (and accumulation) sites of the
type members of both families, with available data indicating that in this respect
other viroids behave like their corresponding type species. Within each genus, the
criteria to demarcate viroid species are an arbitrary level of below 90% sequence
similarity and distinct biological properties. Viroids, like viruses, propagate in their
hosts as populations of closely related sequence variants (quasi-species), although
one or more may predominate in the population. Heat stress may significantly alter
the structure of viroid quasi-species (62). Some viroid variants with minor changes
affecting certain regions are directly related to specific diseases (57, 71, 83) or to
dramatic alterations in symptom severity (21, 93, 103).
BIOLOGY
Host Range and Host Specificity
Viroids are the etiologic agents of a number of diseases affecting economically
important herbaceous and ligneous plants including potato, tomato, cucumber,
hop, coconut, grapevine, several subtropical and temperate fruit trees (avocado,
peach, apple, pear, citrus, and plum), and some ornamentals (chrysanthemum and
coleus). Coconut cadang-cadang viroid (CCCVd) and Coconut tinangaja viroid
(CTiVd) infect monocotyledons, whereas the others infect dicotyledons. Some
viroids, among which the most instructive example is Hop stunt viroid (HSVd),
have wide host ranges but others, exemplified by those forming the family Avsun-
viroidae, are mainly restricted to their natural hosts. A single nucleotide substi-
tution converts PSTVd from noninfectious to infectious for Nicotiana tabacum
(108).
Although most viroids are transmitted mechanically and some through seed
or pollen, with only Tomato planta macho viroid (TPMVd) known to be aphid-
transmissible under specific ecological conditions, the most efficient transmission
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 121
route for viroids is vegetative propagation of infected material. This explains why
certain grapevine and, specifically, citrus cultivars propagated on infected cultivars
or rootstocks contain complex mixtures of different viroids.
Symptoms and Ultrastructural Effects
Some viroids have destructive consequences, as illustrated by CCCVd that has
killed millions of coconut trees in the Philippines, whereas others affect leaves
(chlorosis in spots or covering the whole blade, epinasty, rugosity, and necro-
sis), stems (pitting, internode shortening, and dwarfing), bark (scaling, cracking,
cankers), flowers (size reduction, broken lines on petals), fruits (discolorations and
skin deformations, suture cracking), seeds (enlarged stones), and reserve organs
(malformations), as well as less conspicuous effects including delays in foliation,
flowering and ripening, and growing pattern (open habit) of mature trees (Figure 2).
Certain viroids only induce symptoms in a particular organ (bark or fruit), whereas
others have more general effects. Infections caused by a few viroids result in very
mild or no symptoms. Absence of symptoms is common in naturally infected wild
plants, which can act as reservoirs. Symptom expression is generally favored by
high light intensity and, particularly, high temperature. Viruses, by contrast, have a
broader range of environmental conditions for optimal symptom expression. These
differences may explain why viroids mainly affect crops grown in tropical or sub-
tropical areas (and in greenhouses), and also why curing some viroid infections is
recalcitrant to thermotherapy.
Ultrastructural studies on the cytopathic effects induced by members of the
family Pospiviroidae have shown paramural bodies known as plasmalemmasomes
and aberrations of the thylakoid membranes of chloroplasts (94). Parallel stud-
ies with members of the family Avsunviroidae have revealed grossly disorganized
chloroplasts and membranous bodies in the yellow regions of ASBVd-infected av-
ocado leaves, while in completely chlorotic (“bleached”) leaves some chloroplasts
looked similar to proplastids (23). This latter observation has been reproduced in
PLMVd-infected leaves displaying extensive chlorosis (peach calico), which in
the most dramatic instances covers most of the leaf area (Figure 2h) (M.E. Rodio,
A. Destradis, R. Flores & F. Di Serio, unpublished results).
Cross-Protection
The phenomenon of cross-protection refers to observations that the ability of mem-
bers of both families to infect a host may be influenced by previous infections by
other strains of the same or closely related viroid (53, 68). More specifically, when
a plant pre-infected with a mild viroid strain is challenge-inoculated with a se-
vere strain of the same viroid, the typical symptoms of the second strain and the
accumulation level of its RNA are blocked or attenuated for a certain time. On
this basis, even before PSTVd, PLMVd, and CChMVd were identified as viroids,
the existence of mild or nonsymptomatic strains thereof was postulated and used
to develop cross-protection bioassays. It was also suggested that the nature of
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
122 FLORES ET AL.
CChMVd was uncharacteristic (68). The mechanisms underlying cross-protection
have not been determined but may be related to RNA silencing (see below).
REPLICATION
Circular and Oligomeric RNA Templates: Support
foraRolling-Circle Mechanism
On finding that in PSTVd-infected tomato the most abundant viroid circular RNA,
arbitrarily considered as having (+) polarity, is accompanied by oligomeric ()
RNAs, it was proposed that the latter were replicative intermediates resulting from
reiterative transcription of the former (6). The existence of CEVd ()RNAse-
quences in infected Gynura aurantiaca (40) had already indicated that viroid
replication was an RNA-based process and undermined the idea of participa-
tion of DNA intermediates. Moreover, differential centrifugation studies showing
that PSTVd (24) and its complementary strands (99) accumulate in the nucleus
strongly suggested the involvement of a nuclear RNA polymerase in replication.
Re-examination of this question with PSTVd and other members of its family using
finer approaches (in situ hybridization and confocal laser scanning and transmis-
sion electron microscopy) confirmed (5, 44) and extended these observations (81).
In contrast, parallel experiments first with ASBVd (4, 56, 63) and then with
PLMVd (9) showed the preferential accumulation of their (+) and () strands in the
chloroplast, indicating involvement of the enzymatic machinery of this organelle
in the replication of members of the family Avsunviroidae.
Rolling-Circle Mechanism: Asymmetric and
Symmetric Pathways
Figure 3 displays the two alternative pathways: the asymmetric, first proposed to
account for replication of PSTVd (6, 7, 31) and HSVd (48), and the symmetric,
initially advanced on theoretical grounds (6) and then experimentally supported
for ASBVd (18, 46) and PLMVd (9). Other members of both families appear to
replicate following the same pathway as their type species. The difference discrim-
inating both pathways is the () template: The monomeric () circular RNA has
been detected in ASBVd-infected avocado (18, 46) and PLMVd-infected peach
(9), but not in PSTVd-infected tissues (6, 7, 31), or in electroporated protoplasts
(79) in which oligomeric ()strands accumulate. Thus, cleavage and ligation oc-
cur in (+) and () strands in the symmetric pathway with two rolling circles, but
only in (+) strands in the asymmetric pathway with a single rolling circle.
The three catalytic activities required—RNA polymerase, RNase, and RNA
ligase (Figure 3)—were initially presumed to reside in host proteins. However,
in the family Avsunviroidae, the second activity is now known to be mediated
by hammerhead ribozymes embedded in both polarity strands, a finding with far-
reaching implications. In the context of the first activity, the question of how
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 123
viroids redirect the template specificity of certain host DNA-dependent RNA
polymerases to transcribe RNA is one of the most interesting challenges
unresolved.
RNA Polymerization, Cleavage, and Ligation:
Role of Enzymes and Ribozymes
In light of recent reviews (34, 100), here we summarize only the main features,
novel data, and some unresolved issues regarding RNA polymerization, cleav-
age, and ligation. In vivo and in vitro transcription analyses in the presence of
α-amanitin indicate that RNA strand elongation in representative members of the
family Pospiviroidae is abolished by nanomolar concentrations of this fungal toxin,
which typically inhibit the nucleoplasmic RNA polymerase II (32, 64, 91). In line
with this view, a chromatin-enriched fraction from CEVd-infected tomato, puri-
fied by affinity with a monoclonal antibody to the carboxy-terminal domain of
the largest subunit of RNA polymerase II, has been shown to contain the viroid
(+) and () strands (106). Similar analyses in the presence of tagetitoxin support
the role of a nuclear-encoded chloroplastic RNA polymerase (NEP), or another
polymerase resistant to this bacterial inhibitor, in ASBVd strand elongation (67).
In vitro studies with PLMVd and RNA polymerase of Escherichia coli suggest
the participation of the eubacterial-like plastid-encoded polymerase (PEP) (75),
although this is a nonphysiological system. Moreover, synthesis (and accumula-
tion) of PLMVd is particularly active in areas exhibiting peach calico, in which
development of proplastids into chloroplasts and processing of chloroplastic rRNA
precursors (and, consequently, translation of plastid-encoded proteins) appear to
be impaired (M.E. Rodio, R. Flores & F. Di Serio, unpublished results). These lat-
ter observations are more consistent with the involvement of a NEP-like enzyme
in PLMVd replication and suggest that other chloroplastic factors mediating this
process (see below) are also nuclear-encoded.
To determine whether initiation of viroid RNAs is site-specific (promoter-
driven) or occurs at random (also allowing complete transcription of the circular
template), data obtained by in vitro capping with [α32P]GTP and guanylyltrans-
ferase, which specifically labels the free 5-triphosphate group characteristic of
chloroplastic primary transcripts, and RNase protection assays have mapped the
initiation of ASBVd (+) and ()RNAs isolated from infected avocado at similar
A+U-rich terminal loops in their predicted quasi-rod-like secondary structures
(66). Primer-extension analysis of the 5termini of PLMVd subgenomic RNAs
from infected peach, presumed to be replication by-products, and in vitro tran-
scription studies with truncated PLMVd RNAs and RNA polymerase of E. coli
suggest that the initiation sites of this viroid also map at terminal loops (75). How-
ever, in vitro capping studies to map the 5-triphosphate termini of PLMVd (+)
and () linear RNAs isolated from infected tissue indicate alternative initiation
sites (S. Delgado, A.E. Mart´ınez de Alba, C. Hern´andez & R. Flores, unpublished
results). Still controversial are the initiation sites in the family of nuclear viroids
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
124 FLORES ET AL.
(Pospiviroidae), an issue that has been addressed by in vitro transcription of the
PSTVd monomeric (+) circular RNA with a nuclear extract from potato or with
purified RNA polymerase II from wheat germ and tomato (100).
Correct processing to the monomeric (+) circular forms has been demon-
strated through assays in vitro with potato nuclear extracts and longer-than-unit
(+) PSTVd RNAs (2, 102) and in vivo with Arabidopsis thaliana transformed
with cDNAs expressing dimeric (+) transcripts of five representative members
of the family Pospiviroidae (17). These results indicate that the RNase activity
catalyzing cleavage of longer-than-unit (+) strands is a host enzyme(s) whose
site-specificity is determined by a particular RNA folding. Specifically, cleavage
of PSTVd (+) strands is proposed to be driven by a branched structure with a
GNRA tetraloop, which subsequently switches to an extended conformation with
an E loop promoting ligation (2). However, this mechanism may not apply to other
members of the family Pospiviroidae that are unable to form the GNRA tetraloop
and the loop E. On the other hand, it has been found that a dimeric () transcript
of HSVd expressed transgenically in A. thaliana fails to be processed (17) and
that in infected cultured cells and plants PSTVd () strands accumulate in the
nucleoplasm, whereas the (+) strands are localized in the nucleolus as well as in
nucleoplasm (Figure 3) (see below). Such findings suggest that viroid (+) strands
are processed in the nucleolus where processing of the rRNA and tRNA precursors
also occurs. Which factors determine this differential traffic of viroid (+) and ()
strands remain an intriguing issue.
In members of the family Avsunviroidae,however, the RNase activity is due not
to an enzyme but to a hammerhead ribozyme, a small RNA motif embedded in both
polarity strands that, through a transesterification in the presence of Mg2+, self-
cleaves at a specific phosphodiester bond producing 5-OH and 2,3-cyclic phos-
phodiester termini (47, 78; reviewed in 35). Hammerhead ribozymes are formed
by a central core of conserved sequences flanked by three double-stranded regions
with loose sequence requirements that are capped by loops. X-ray crystallography
has revealed a complex array of non-canonical interactions between the nucleotides
of the central core, explaining why they are strictly conserved in natural ribozymes
and illustrating that the actual shape does not resemble a hammerhead but rather
an inverted Y in which the stems III and II are almost co-linear (Figure 1b, inset).
There is firm evidence supporting the proposal that hammerheads mediate the in
vivo self-cleavage of the multimeric viroid RNAs wherein they are inserted, and
that the reaction is under strict control through a conformational switch between
the active ribozyme, which is transiently formed during replication, and an alter-
native folding that promotes ligation (34). In contrast to the accepted view, recent
data show that modifications of loops 1 and 2 of natural hammerheads induce a
severe reduction in their catalytic activity, indicating that these peripheral regions
play a critical role in catalysis through tertiary interactions between some of their
nucleotides that may favor the active site at the low magnesium concentration
existing in vivo (20, 54) (Figure 1b, inset). These interactions could be stabilized
by chloroplastic proteins behaving as RNA chaperones, which would explain why
they facilitate the hammerhead-mediated self-cleavage of a viroid RNA (16).
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 125
Data on the third replication step are limited, but in PSTVd and nuclear viroids
the reaction is presumably catalyzed by a host enzyme similar to the wheat germ
RNA ligase. Support for this hypothesis comes from data in vitro showing that this
enzyme mediates circularization of the monomeric linear PSTVd RNA isolated
from infected tissue (8) and that a potato nuclear extract can process longer-
than-unit PSTVd RNAs to the monomeric circular forms (2, 102), and in vivo
indicating that A. thaliana seems to have a similar enzyme able to circularize the
monomeric linear forms of representative members of the family Pospiviroidae
(17). Still unclear is whether a chloroplastic RNA ligase exists for members of
the family Avsunviroidae or, alternatively, the reaction is autocatalytic. Support
for this latter view is based on the in vitro self-ligation of the monomeric linear
PLMVd RNAs resulting from hammerhead-mediated self-cleavage, which mostly
leads to 2,5- instead of the conventional 3,5-phosphodiester bonds (14) and the
proposal that these atypical bonds are present in circular PLMVd RNAs isolated
from infected tissue and impede their in vitro self-cleavage (15). In line with these
results, early in vitro studies with RNA of human hepatitis delta virus (HDV)
which, like viroids, also replicates through a rolling-circle mechanism mediated
by ribozymes (albeit of a non-hammerhead class), showed ligation of ribozyme-
cleaved sequences in protein-free conditions (reviewed in 101). However, these
conditions were far from physiological, and more recent studies have shown that a
host-specific function is needed for ligation of the RNAs resulting from cleavage
by a wide variety of ribozymes, which include hammerhead and HDV ribozymes
(84). Furthermore, the hammerhead-generated 5-OH and 2,3-cyclic phospho-
diester termini are those typically required by a wheat germ-like RNA ligase, a
feasible alternative to self-ligation and, in contrast to a previous report (15), the
circular PLMVd RNAs isolated from infected tissue appear to self-cleave in vitro
(S. Delgado, C. Hern´andez & R. Flores, unpublished results).
MOVEMENT
Viroid movement received scant attention in early studies on these peculiar patho-
gens but an increasing number of recent reports are now elucidating the mecha-
nisms and putative host factors involved in this step of the viroid infectious cycle.
This area has been further promoted by the possible relevance of these results to
the general processes of RNA transport in plants, including mRNAs and mobile
silencing signals (see below). In contrast to viruses, which encode their own move-
ment proteins, viroids must interact directly with host factors for mobility. Viroids
therefore offer a unique system to dissect RNA traffic in plants.
Intracellular Movement
After entering a cell, the viroid RNA must move to its replication site, either the
nucleus or the chloroplast, to generate the progeny for release to the cytoplasm
to invade neighboring cells. Import of fluorescein-labeled PSTVd transcripts into
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
126 FLORES ET AL.
the nucleus has been studied in permeabilized tobacco cells and shown to be a
cytoskeleton-independent process mediated by a specific and saturable receptor
(109). This process was demonstrated to be sequence or structure specific because
no import was observed for mRNA fragments of the same size or for twoviroids that
replicate and accumulate in the chloroplast. These results were validated in planta
with an RNA virus vector expressing a green fluorescent protein (GFP) reporter
gene bearing an intron that was either unmodified or contained an embedded full-
length PSTVd copy. Inoculation with these cytoplasmic replicating viral constructs
caused fluorescence only with the PSTVd-containing vector, indicating that the
viroid sequence targeted the recombinant RNA to the nucleus where the intron
was removed, with the spliced mRNA returning to the cytoplasm where it was
translated (110).
Overall, these data suggest that nuclear import of PSTVd results from interac-
tion of a viroid sequence or structural motif with cellular factors that may lead to
the formation of a ribonucleoprotein complex, which would shuttle the viroid to the
nucleus. Pertinent to this point is the identification of a bromodomain-containing
protein from tomato that binds specifically PSTVd in vitro and in vivo (61). This
Viroid RNA-binding Protein 1 (VirP1), which was isolated from an RNA-ligand
screening through its ability to interact with PSTVd (+)RNA,isamember of
afamily of transcriptional regulators associated with chromatin remodeling. As
expected, VirP1 contains a nuclear localization signal and is a candidate to mediate
PSTVd transfer (and that of other related viroids) into or out of the nucleus. Further
work has mapped the RNA motif responsible for the specific interaction between
VirP1 and PSTVd to an asymmetrical internal loop (termed RY motif owing to its
base composition), within the right terminal domain, which is present twice in this
viroid and in most members of the genus Pospiviroid, once in HSVd, and is partially
conserved in the genus Cocadviroid (39, 58). Mutations in any of the two RY motifs
of PSTVd abolished infectivity, a finding that supports its biological relevance.
The general picture of the intracellular viroid movement becomes more complex
when analyzed at the organelle level. Using improved sample preparation and in
situ hybridization, it has recently been shown that the PSTVd () strand localizes
in the nucleoplasm of infected tomato and N. benthamiana plants or cultured cells,
whereas the (+) strand localizes in the nucleolus as well as in the nucleoplasm
(Figure 3), with distinct spatial patterns that may represent successive stages of the
viroid RNA migration or processing (81). These results suggest that after synthesis
of PSTVd (+) and ()RNAsinthe nucleoplasm, the latter stays anchored to this
compartment while the (+) strand is transported to the nucleolus. This finding
points to the existence of highly specialized machinery in eukaryotic cells able to
discriminate the opposite strands of an RNA and may have implications in gene
regulation and pathogen infection.
As regards intracellular movement in members of the family Avsunviridae, the
transfer mechanism to the chloroplast has yet to be investigated. Novel transport
pathways may well be discovered in plant cells, given that no other alien or cellular
RNAs have been reported to traffic inside this organelle.
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 127
Cell-to-Cell Movement
Once it has infected the first cells, a viroid has to colonize adjacent cells prior to
invading the more distal plant parts. Proteins and nucleic acids, either endogenous
or of viral origin, have been reported to move cell-to-cell via the plasmodesmata
(PD), the plant organella providing cytoplasmic connections between cells. Stud-
ies with PSTVd suggest that viroids follow the same pathway for intercellular
movement (28). PSTVd was retained when microinjected into symplasmically
isolated guard cells of mature tomato and tobacco leaves; however, when injected
into symplasmically connected mesophyll cells, the viroid moved quickly from
cell to cell. Remarkably, the fusion to PSTVd aided the transport through PD of
an otherwise nonmobile RNA, suggesting that the viroid contains a motif for PD
transport.
Long-Distance Movement
Systemic spread of viroids occurs through the vasculature and allows their access
to tissues far away from those initially infected. As in viruses, this transport in-
volves loading the viroid into the phloem and follows the flow of photoassimilates
from the photosynthetic source to sink tissues/organs of the plant (72, 111). Evi-
dence of this movement at the cellular level has been provided for PSTVd by in situ
hybridization, which also showed that trafficking within the phloem is probably
sustained by viroid replication and tightly regulated by plant developmental and
cellular factors (111). The viroid was not detected in shoot apical meristems (SAM)
of infected N. benthamiana or tomato plants, whereas it was present in the vascu-
lar tissues (most probably the procambium and/or protophloem) below the SAM,
suggesting that PD at some cellular boundaries between the SAM and the rest of
plant body restrict PSTVd trafficking into the SAM. Moreover, PSTVd was absent
in developing flowers, whereas mature flowers contained the viroid in parenchyma
cells of sepals but, intriguingly, not in petals, stamens, styles, or ovaries, although
the phloem connections were already established. The absence of PSTVd in some
floral organs of N. benthamiana may be attributable to restricted traffic in these
organs and not to replication suppression (112). This restriction could be nonop-
erative under certain conditions because PSTVd is seed-transmissible in tomato
and therefore able to infect ovules and/or pollen eventually. In contrast to the
strong vascular tropism that PSTVd shows below the SAM in tomato, PLMVd
appears able to invade cell layers very close to the SAM in peach (M.E. Rodio, S.
Delgado, M.D. G´omez, R. Flores & F. Di Serio, unpublished results), suggesting
that different mechanisms regulate movement in the two viroid families.
Viroid translocation via the phloem is most probably facilitated by host pro-
teins. Evidence has been obtained in vitro and in vivo of the formation of a ri-
bonucleoprotein complex between HSVd and one of the most abundant phloem
polypeptides of cucumber, the dimeric lectin known as Phloem Protein 2 (PP2)
(37, 38, 69). This protein fulfils the requirements of an RNA chaperone protein in-
volved in systemic movement of HSVd: RNA-binding activity probably mediated
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
128 FLORES ET AL.
by a double-stranded RNA binding motif, ability to interact with PD and increase
their exclusion limit, and competence for long-distance translocation. Specifically,
this protein is able to move and aid HSVd movement through intergeneric grafts
involving a host rootstock and a nonhost scion, although the viroid remains largely
confined within the vascular tissue of the scion, suggesting that additional host
factors are required for HSVd to leave the phloem efficiently (38). Indeed, phloem
entry and exit are probably mediated by different motifs, as revealed by the iden-
tification of two PSTVd mutants able to enter and replicate in the phloem of N.
tabacum but unable to leave the vascular tissue (112). The same viroid variants
infected all cells of the upper uninoculated leaves in N. benthamiana, indicating
that the regulation mechanisms of phloem-mediated RNA traffic may differ in
distinct plant species. Moreover, in PSTVd, a motif has recently been identified
that mediates trafficking from the bundle sheath into the mesophyll but not in the
opposite direction (82).
It has been suggested that VirP1 from tomato, like PP2 from cucumber, is
involved in the systemic spread of PSTVd in this host, given that a PSTVd mutant
defective for VirP1 binding is unable to spread systemically (58). However, in
contrast to PP2, the ability of VirP1 to interact with PD and move long distance
has yet to be proved. In conclusion, viroids must have evolved to exploit existing
transport routes and, given their peculiar characteristics, one may expect these
pathogens to be instrumental in elucidating structural traits, mechanisms, and
cellular factors that mediate RNA traffic in plants.
PATHOGENESIS
Lacking protein-coding capacity, viroids must incite disease by direct interaction
of their genomic RNA or derivatives thereof with host factors (proteins or nucleic
acids). This primary interaction triggers a cascade of events, still poorly understood,
that eventually lead to macroscopic symptoms.
Is the Mature Viroid RNA the Direct Pathogenic Effector?
In PSTVd, mutations of 3–4 nucleotides, with marked effects on symptoms, have
been mapped at a virulence-modulating (VM) region that overlaps a premelting
(PM) region within a domain of the rod-like structure (93). Because a similar
situation was observed in CEVd (103), this domain was termed pathogenic (P)
(52) (Figure 1a). Although an inverse correlation was found between the ther-
modynamic stability of the PM region and virulence in tomato for some PSTVd
strains (93), other data do not support this correlation (70). Alternatively, from
comparisons of the most stable secondary structures of PSTVd variants of differ-
ent pathogenicity and considerations about bending of RNA helices, it has been
advanced that major differences in their VM region geometry and concomitant al-
terations in RNA-protein interactions are the primary cause of viroid pathogenicity
(70, 92). In this context, a correlation between the virulence of PSTVd strains and
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 129
the activation of certain protein kinases has been reported (reviewed in 26). It is also
possible that differential interactions with host proteins involved in viroid replica-
tion, movement, or accumulation could be the initial event in viroid pathogenesis
(Figure 4). Moreover, other data indicate the contribution of additional domains to
symptom expression (89), including the C domain, in which a specific nucleotide
(A257) in PSTVd induces severe growth stunting and flat top of the tomato shoot
(80). Pathogenicity determinants outside the P domain have also been identified in
other members of the family Pospiviroidae such as CCCVd (85) and HSVd (71,
83).
Within the family Avsunviroidae, motifs involved in pathogenesis have also
been recognized. ASBVd variants slightly diverging in the poly-A right terminal
loop have been associated with different leaf symptoms (95), although difficulties
inherent to bioassays in avocado have precluded a direct analysis. In CChMVd,
a tetraloop in the in vivo branched RNA conformation has been mapped as the
major pathogenicity determinant because site-directed mutagenesis and bioassays
have shown that the change from UUUC to GAAA converts a variant from se-
vere into latent without altering the final viroid titer (19, 21). A similar method-
ology was used to demonstrate that PLMVd variants containing a 12–13-nt in-
sertion incite peach calico. This insertion is always found in the same position,
has limited sequence variability, and folds itself into a hairpin (57). Moreover,
the pathogenicity determinant might be restricted to the U-rich tetraloop of this
hairpin. Although the precise role of this loop in symptom development is not
known, electron microscopy observations have revealed that chloroplast differ-
entiation is blocked in symptomatic tissues (M.E. Rodio, S. Delgado, R. Flores
&F.DiSerio, unpublished results). The identification of pathogenicity determi-
nants with a similar structure (a U-rich tetraloop) in CChMVd and PLMVd is
intriguing.
Instead of proteins as the primary host target, base-pair interactions between
PSTVd and host RNAs, resulting in interference with rRNA maturation, mRNA
splicing, or 7S RNA assembly into the signal recognition particle, were proposed
as possible molecular events initiating pathogenesis (reviewed in 26). Although
these hypotheses fail to explain the differential symptoms induced by the same
viroid variant on closely related species (26), the observation that the PSTVd (+)
strand is able to migrate to the nucleolus and cause a redistribution of the small
nucleolar RNAs has led to the suggestion that PSTVd could incite deleterious
effects on rRNA processing and related events by competing for certain nucleolar
factors (81).
Early studies on the signal-transduction cascade leading to symptom expression
showed that viroids activate a general plant defense system based on pathogenesis-
related (PR) proteins. This system is also induced by other pathogens including
viruses, bacteria and fungi, which modify the concentrations of certain hormones
and metabolites such as polyamines (13). More recently, a macroarray-based ap-
proach has shown that PSTVd infection alters the gene expression pattern in tomato
and that the regulation of specific genes, which are not involved in the host reaction
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
130 FLORES ET AL.
to other pathogens, may be affected in a strain-dependent manner (50). The role
of these genes in pathogenesis is unknown, except in tomato plants infected with
a PSTVd strain with the U257A substitution; here the stunted growth is caused
by restricted cell expansion, but not cell division or differentiation, and has been
correlated with the down-regulated expression of an expansin gene, LeExp2 (80).
Although it is generally accepted that viroid diseases are induced by specific
interference with regulation of host gene expression, many questions about the
underlying mechanism remain unanswered. The discovery that a group of small
antisense RNAs, namely the small interfering RNAs (siRNAs) and the microRNAs
(miRNAs) (reviewed in 1, 11), mediate a regulatory layer of host gene expression
in eukaryotes has led to new attractive hypotheses for viroid pathogenesis.
Do Viroids Exert their Pathogenic Effects via Specific siRNAs?
miRNAs (21–24 nt long) negatively regulate the expression of genes involved
mainly in development (11). They are generated by an enzyme of the RNase
III class (Dicer) (3), Dicer-like (DCL) in plants, acting on miRNA precursors,
which are endogenous nonprotein-coding transcripts adopting a hairpin conforma-
tion with partially double-stranded regions. Depending on the degree of sequence
complementarity with their cognate mRNAs, miRNAs guide the RNA-induced
silencing complex (RISC) (43) to degrade the target RNA or, less frequently in
plants, to repress its translation. Plant siRNAs (21–26-nt long) are associated with
both transcriptional gene silencing (TGS) and posttranscriptional gene silencing
(PTGS) (reviewed in 1). TGS controls gene expression negatively by siRNA-
directed DNA methylation of certain promoters, a process that is linked to histone
modification in plants. PTGS, a sequence-specific RNA-degradation mechanism,
was first identified in plants as a defense response against invading RNAs from
transgenes, transposons, and viruses (reviewed in 1, 55). siRNAs play a key role
in this system because, like miRNAs, they are loaded into RISC and guide it to
degrade specific RNAs. siRNAs also resemble miRNAs in that they are generated
by Dicer, although acting on double-stranded RNA (dsRNA) in this case. Most
transgenic and transposon dsRNAs derive from transcription of inverted repeats in
the nucleus, with other dsRNAs that accumulate in the cytoplasm resulting from
replication of single-stranded RNA (ssRNA) viruses. To counteract this antiviral
defense, plant viruses code for specific proteins, called silencing suppressors, that
inhibit PTGS (reviewed in 55).
The possibility that viroids may influence host gene expression at both post-
transcriptional and transcriptional levels and, indirectly, induce symptoms (see
below) is supported by the recent identification of viroid-specific siRNAs in plants
infected by members of the families Pospiviroidae (49, 59, 73) and Avsunviroidae
(59, 60), together with the previous finding that replicating PSTVd induces de
novo methylation of PSTVd sequences transgenically inserted in the plant genome
(107). Where and how these viroid-specific siRNAs are generated is unresolved.
Since DCL isoenzymes with nuclear localization have been described (reviewed
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 131
in 1) (Figure 4), a nuclear origin for the siRNAs in the family Pospiviroidae
seems plausible. In contrast, because no chloroplastic DCL has been reported,
siRNA generation in the cytoplasm during viroid cell-to-cell movement seems
more likely in the family Avsunviroidae. This possibility could also apply to nu-
clear viroids, given that PSTVd-derived siRNAs accumulate in the cytoplasm (22)
(Figure 4). Regarding the templates, the primary candidates for siRNA genera-
tion in the family Pospiviroidae are the nuclear dsRNA intermediates of viroid
replication. However, the mature genomic forms (structurally similar to miRNAs
precursors) could also be degraded by DCL-1 in the nucleus or, alternatively,
serve as templates for the synthesis of secondary dsRNAs catalyzed by a cytoplas-
mic RNA-dependent RNA polymerase (RdRp) (90) (Figure 4). Indeed, although
PSTVd and ASBVd monomeric forms are resistant in vitro to human Dicer (12),
they appear to be sensitive to wheat germ DCL (A.E. Mart´ınez de Alba and R.
Flores, unpublished results). In this framework, direct characterization of the small
RNAs derived from replicating PSTVd indicates that they are predominantly of the
(+) polarity resembling, as a population, miRNAs rather than siRNAs (B. Ding,
personal communication).
Are viroid-specific siRNAs relevant in pathogenesis or are they just by-products
of the RNA silencing machinery? The in vivo concentration of these siRNAs cannot
explain differences in symptom development because their levels are essentially
the same in infections induced by severe, mild, or latent strains of PSTVd and
CChMVd (60, 73). However, an inverse correlation exists in chloroplastic viroids
between the accumulation of genomic viroid forms and their corresponding siR-
NAs: the in vivo PLMVd and CChMVd concentrations are low but their siRNAs
are easily detectable, whereas in tissues accumulating a very high ASBVd con-
centration the corresponding siRNAs are undetectable (60) or accumulate poorly
(59). This finding is consistent with the siRNA participation in a PTGS defense
response of certain hosts aimed at attenuating the in vivo viroid titer (60); this re-
sponse should take place in the cytoplasm where RISC is assumed to act. Other data
support this view indirectly. First, PSTVd infection of transgenic plants, express-
ing a reporter gene 3fused to partial-length PSTVd sequences, activates degra-
dation of the recombinant transcript (and methylation of the nuclear transgenic
viroid DNA) (104), probably via the siRNAs that result from PSTVd replication.
Second, preliminary observations indicate that the infectivity of some viroids is
reduced when they are coinoculated with an excess of their cognate dsRNAs (A.
Carbonell, S. Gago & R. Flores, unpublished results), suggesting that the latter
are processed by DCL in vivo and the resulting siRNAs target the genomic forms
for RISC-mediated degradation (Figure 4). And third, accumulation of PSTVd-
specific siRNAs precedes tomato plant recovery from severe symptoms of PSTVd
infection (88). These results argue against the proposal that viroids are resistant
to RNA silencing–mediated degradation and might have evolved their compact
secondary structure to escape this host defense pathway (105). Such folding might
also have appeared to provide resistance against RNases or certain inactivating
proteins that target RNA single-stranded regions (74).
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
132 FLORES ET AL.
However, some viroid-specific siRNAs might also direct host DNA methylation
(76) or act like endogenous miRNAs targeting host mRNAs for degradation (59, 73,
105) (Figure 4). Supporting this idea, transgenic tomato plants expressing inverted
repeats of an almost complete (noninfectious) PSTVd sequence showed symptoms
similar to, though milder than, those induced by infectious PSTVd RNA. These
transgenic plants accumulated PSTVd-derived siRNAs, whereas neither siRNAs
nor typical symptoms were detected in plants expressing direct repeats of the same
viroid sequence (105). However, the effects of expressing sequences from a mild
PSTVd strain were not determined, very few small RNAs derived from replicating
PSTVd correspond to domains responsible for symptom expressiod(same)]TJT*[(viroid)-262.5(sequence)-262.4((105).)-262.5(Ho)25(we)25(v)15(e)0(r)40(,)-262.4(the)-262.5(ef)25(fects)-262.5(of)-262.4(e)15(xpressing)-262.5(sequences)-262.4(from)-262.5(a)-262.4(mild)]TJT*[(PSTVd)-201.4(strain)-201.4(were)-201.3(not)-201.4(determined,)-201.2(v)15(ery)-201.4(fe)25(w)-201.4(small)-201.3(RN)35(As)-201.4(deri)25(v)15(e)0(d)-201.4(from)-201.3(replicating)]TJT*[(PSTVd)-204.9(correspond)-204.8(to)-204.9(domains)-204.9(responsible)-204.8(for)-204.9(symptom)-204.9(e)15(xpressiod(same)]TJT*[(sam2TJT*[(PSTVd)-204.9(h24not)-201.h)]xpres19 Tm4.8(orssam2TJTw75.01.4(determined,)-2-r.1 -170 9.9626343)-20athw0.2 ays26343)-2 7116(tar343)2(Inasmuch26343)-2as26343 P43 s43ers43 not ho01.ol(v)15(ercomplete)-TG5219 Tm4.8(orssa259(,)4(der521ader521co)0(drder521defen32)T5219s)-t.1 -17yed fr5215.1(v521T*[(P(v5210.5649 der521dromoder521(ef)25(.8(to)-204.op2)T5TJTw75.0T5TJT9(domains)5TJ3strain)-20?0T5TJTP-201ing)]TJ5TJ3sco)ld0T5TJTb2dordedTJ5TJ3s[(PST5TJTpreirPST5TJT)15( -17.5TJ3sconhe)-han, 000117.933m3 ref245es1-149.6(,f245e204)19(,)utf245es1one v6(,f245e 755.9245e 7b)-2245e20 rtressi ref245et 017.933o.29995.8(i8(ocii ref995.1Tw75.0T995.8()-201.hs.0T995.8(262erni r(e)15(xpree)-65(,f995.1Tpresxpry0T995.8(co)ld0T995.8(me)]TJT*0T995.8(w75.0T995.8(pressi95.1T1.8(acc95.8(meetected)-19bor.3(re2.7)-3ed)e3(re2.7)sy)-2m3(re2.7)because3(re2.7)(endogenre2.7)-s)-201.4enre2.7)has263e2.7)bee4enre2.7)i8(ocii egenre2.7)w75.0T9e2.7)-sild
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 133
of life”). Additionally, it has led to discoveries transcending virology and the
plant world: the hammerhead ribozyme, with deep functional, biotechnological,
and evolutionary repercussions (47, 78; reviewed in 35), and RNA-dependent
DNA methylation (107) that mediates transcriptional silencing. This illustrates,
once again, how unveiling apparently marginal details of a specific system can
illuminate far-reaching questions. But this may not be the end. Recently, viroids
have become a tool for dissecting aspects of transcriptional and posttranscriptional
gene silencing, as well as for understanding how minimal RNAs can elicit disease
(reviewed in 26, 34, 100). Moreover, viroids are being used as probes to study
the factors governing the intracellular, cell-to-cell, and long-distance movement
of RNAs (81, 109, 111, 112). Within this context, studies on chloroplast viroids
could elucidate how a foreign RNA manages to cross the membrane surrounding
this organelle. Last but not least, whether viroids exist out of the plant kingdom
is still a challenging and exciting question. After all, since viruses were initially
found in plants and subsequently in most types of other organisms, why should
viroids be restricted to the green world?
ACKNOWLEDGMENTS
We apologize to colleagues whose articles were not cited because of space lim-
itations. We thank B. Ding and R.A. Owens for access to unpublished data and
suggestions, respectively. Work in R. Flores and J.A. Dar`os laboratories has been
supported by the Ministerio de Ciencia y Tecnolog´ıa (BMC2002-03,694) and the
Generalidad Valenciana (Spain).
The Annual Review of Phytopathology is online at
http://phyto.annualreviews.org
LITERATURE CITED
1. Baulcombe D. 2004. RNA silencing in
plants. Nature 431:356–63
2. Baumstark T, Schr¨oder ARW, Riesner
D. 1997. Viroid processing: switch from
cleavage to ligation is driven by a change
from a tetraloop to a loop E conformation.
EMBO J. 16:599–610
3. Bernstein E, Caudy AA, Hammond SM,
Hannon GJ. 2001. Role for a biden-
tate ribonuclease in the initiation step
of RNA interference. Nature 409:363–
66
4. Bonfiglioli RG, McFadden GI, Symons
RH. 1994. In situ hybridization local-
izes avocado sunblotch viroid on chloro-
plast thylakoid membranes and coconut
cadang-cadang viroid in the nucleus.
Plant J. 6:99–103
5. Bonfiglioli RG, Webb DR, Symons RH.
1996. Tissue and intra-cellular distribu-
tion of coconut cadang-cadang viroid
and citrus exocortis viroid determined by
in situ hybridization and confocal laser
scanning and transmission electron mi-
croscopy. Plant J. 9:457–65
6. Branch AD, Robertson HD. 1984. A
replication cycle for viroids and other
small infectious RNAs. Science 223:450–
54
7. Branch AD, Benenfeld BJ, Robertson
HD. 1988. Evidence for a single rolling
circle in the replication of potato spindle
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
134 FLORES ET AL.
tuber viroid. Proc. Natl. Acad. Sci. USA
85:9128–32
8. Branch AD, Robertson HD, Greer C,
Gegenheimer P, Peebles C, Abelson J.
1982. Cell-free circularization of viroid
progeny RNA by an RNA ligase from
wheat germ. Science 217:1147–49
9. Bussi`ere F, Lehoux J, Thompson DA,
Skrzeczkowski LJ, Perreault J-P. 1999.
Subcellular localization and rolling circle
replication of peach latent mosaic viroid:
hallmarks of group A viroids. J. Virol.
73:6353–60
10. Bussi`ere F, Ouellet J, C ˆot ´eF,L´evesque
D, Perreault JP. 2000. Mapping in solu-
tion shows the peach latent mosaic viroid
to possess a new pseudoknot in a com-
plex, branched secondary structure. J. Vi-
rol. 74:2647–54
11. Carrington JC, Ambros V. 2003. Role of
microRNAs in plant and animal develop-
ment. Science 301:336–38
12. Chang J, Provost P, Taylor JM. 2003. Re-
sistance of human hepatitis delta virus
RNAs to dicer activity. J. Virol. 77:11910–
17
13. Conejero V, Bell´es JM, Garc´ıa-Breijo F,
Garro R, Hern´andez-Yago J, et al. 1990.
Signalling in viroid pathogenesis. In
Recognition and Response in Plant-Virus
Interactions, ed. RSS Fraser, NATO ASI
Series, H 41:233–61. Berlin/Heidelberg:
Springer-Verlag
14. Cˆot´eF,Perreault JP. 1997. Peach la-
tent mosaic viroid is locked by a 2,5-
phosphodiester bond produced by in vitro
self-ligation. J. Mol. Biol. 273:533–43
15. Cˆot´eF,L´evesque D, Perreault JP. 2001.
Natural 2,5-phosphodiester bonds found
at the ligation sites of peach latent mosaic
viroid. J. Virol. 75:19–25
16. Dar`os JA, Flores R. 2002. A chloroplast
protein binds a viroid RNA in vivo and
facilitates its hammerhead-mediated self-
cleavage. EMBO J. 21:749–59
17. Dar`os JA, Flores R. 2004. Arabidopsis
thaliana has the enzymatic machinery for
replicating representative viroid species
of the family Pospiviroidae.Proc. Natl.
Acad. Sci. USA 101:6792–97
18. Dar`os JA, Marcos JF, Hern´andez C, Flo-
res R. 1994. Replication of avocado sun-
blotch viroid: evidence for a symmetric
pathway with two rolling circles and ham-
merhead ribozyme processing. Proc. Natl.
Acad. Sci. USA 91:12813–17
19. De la Pe˜na M, Flores R. 2002. Chrysan-
themum chlorotic mottle viroid RNA: dis-
section of the pathogenicity determinant
and comparative fitness of symptomatic
and non-symptomatic variants. J. Mol.
Biol. 321:411–21
20. De la Pe˜na M, Gago S, Flores R. 2003.
Peripheral regions of natural hammer-
head ribozymes greatly increase their self-
cleavage activity. EMBO J. 22:5561–70
21. De la Pe˜na M, Navarro B, Flores R.
1999. Mapping the molecular determinat
of pathogenicity in a hammerhead viroid:
a tetraloop within the in vivo branched
RNA conformation. Proc. Natl. Acad. Sci.
USA 96:9960–65
22. Denti MA, Boutla A, Tsagris M, Tabler
M. 2004. Short interfering RNAs specific
for potato spindle tuber viroid are found
in the cytoplasm but not in the nucleus.
Plant J. 37:762–69
23. Desjardins PR. 1987. Avocado sunblotch.
In The Viroids, ed. TO Diener, pp. 299–
313. New York: Plenum
24. Diener TO. 1971. Potato spindle tuber
“virus”: a plant virus with properties of
a free nucleic acid. III. Subcellular loca-
tion of PSTV-RNA and the question of
whether virions exist in extracts or in situ.
Virology 43:75–89
25. Diener TO. 1972. Potato spindle tuber vi-
roid VIII. Correlation of infectivity with
aUV-absorbing component and thermal
denaturation properties of the RNA. Vi-
rology 50:606–09
26. Diener TO. 2001. The viroid: biological
oddity or evolutionary fossil? Adv. Virus
Res. 57:137–84
27. Diener TO, Raymer WB. 1967. Potato
spindle tuber virus: a plant virus with
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 135
properties of a free nucleic acid. Science
158:378–81
28. Ding B, Kwon M-O, Hammond R,
Owens RA. 1997. Cell-to-cell movement
of potato spindle tuber viroid. Plant J.
12:931–36
29. Dingley AJ, Steger G, Esters B, Riesner
D, Grzesiek S. 2003. Structural character-
ization of the 69 nucleotide potato spindle
tuber viroid left-terminal domain by NMR
and thermodynamic analysis. J. Mol. Biol.
334:751–67
30. Fadda Z, Dar`os JA, Fagoaga C, Flores R,
Duran-Vila N. 2003. Eggplant latent vi-
roid (ELVd): candidate type species for
anew genus within family Avsunviroidae
(hammerhead viroids). J. Virol. 77:6528–
32
31. Feldstein PA, Hu Y, Owens RA. 1998.
Precisely full length, circularizable, com-
plementary RNA: an infectious form of
potato spindle tuber viroid. Proc. Natl.
Acad. Sci. USA 95:6560–65
32. Flores R, Semancik JS. 1982. Properties
of a cell-free system for synthesis of cit-
rus exocortis viroid. Proc. Natl. Acad. Sci.
USA 79:6285–88
33. Flores R, Dar`os JA, Hern´andez C. 2000.
The Avsunviroidae family: viroids with
hammerhead ribozymes. Adv. Virus Res.
55:271–323
34. Flores R, Delgado S, Gas ME, Carbonell
A, Molina D, et al. 2004. Viroids: the min-
imal non-coding RNAs with autonomous
replication. FEBS Lett. 567:42–48
35. Flores R, Hern´andez C, De la Pe ˜na M,
Vera A, Dar`os JA. 2001. Hammer-
head ribozyme structure and function in
plant RNA replication. Methods Enzymol.
341:540–52
36. Flores R, Randles JW, Bar-Joseph M,
Owens RA, Diener TO. 2004. Viroidae.
In Virus Taxonomy, Eighth Report of the
International Committee on Taxonomy of
Viruses, ed. CM Fauquet, MA Mayo, J
Maniloff, U Desselberger, AL Ball, pp.
1145–59. London: Elsevier/Academic
37. omez G, Pall´as V. 2001. Identication of
an in vitro ribonucleoprotein complex be-
tween a viroid RNA and a phloem pro-
tein from cucumber plants. Mol. Plant–
Microbe Interact. 14:910–13
38. omez G, Pall´as V. 2004. A long-distance
translocatable phloem protein from cu-
cumber forms a ribonucleoprotein com-
plex in vivo with hop stunt viroid RNA. J.
Virol . 78:10104–10
39. Gozmanova M, Denti MA, Minkov IN,
Tsagris M. 2003. Characterization of the
RNA motif responsible for the specific in-
teraction of potato spindle tuber viroid
RNA (PSTVd) and the tomato protein
Virp1. Nucleic Acids Res. 31:5534–43
40. Grill LK, Semancik JS. 1978. RNA se-
quences complementary to citrus exocor-
tis viroid in nucleic acid preparations from
infected Gynura aurantiaca.Proc. Natl.
Acad. Sci. USA 75:896–900
41. Gross HJ, Domdey H, Lossow C, Jank P,
Raba M, et al. 1978. Nucleotide sequence
and secondary structure of potato spindle
tuber viroid. Nature 273:203–8
42. Hadidi A, Flores R, Randles JW, Seman-
cik JS, eds. 2003. Viroids. Collingwood,
Aust.: CSIRO Publ.
43. Hammond SM, Bernstein E, Beach D,
Hannon GJ. 2000. An RNA-directed nu-
clease mediates post-transcriptional gene
silencing in Drosophila cells. Nature
404:293–96
44. Harders J, Lukacs N, Robert-Nicoud M,
Jovin JM, Riesner D. 1989. Imaging of
viroids in nuclei from tomato leaf tissue
by in situ hybridization and confocal laser
scanning microscopy. EMBO J. 8:3941–
49
45. Hern´andez C, Flores R. 1992. Plus and
minus RNAs of peach latent mosaic vi-
roid self-cleave in vitro via hammerhead
structures. Proc. Natl. Acad. Sci. USA 89:
3711–15
46. Hutchins CJ, Keese P, Visvader JE, Rath-
jen PD, McInnes JL, Symons RH. 1985.
Comparison of multimeric plus and minus
forms of viroids and virusoids. Plant Mol.
Biol. 4:293–304
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
136 FLORES ET AL.
47. Hutchins C, Rathjen PD, Forster AC,
Symons RH. 1986. Self-cleavage of plus
and minus RNA transcripts of avocado
sunblotch viroid. Nucleic Acids Res.
14:3627–40
48. Ishikawa M, Meshi T, Ohno T, Okada Y,
Sano T, et al. 1984. A revised replication
cycle for viroids: the role of longer than
unit RNA in viroid replication. Mol. Gen.
Gen. 196:421–28
49. Itaya A, Folimonov A, Matsuda Y, Nel-
son RS, Ding B. 2001. Potato spindle tu-
ber viroid as inducer of RNA silencing in
infected tomato. Mol. Plant-Microbe In-
teract. 14:1332–34
50. Itaya A, Matsuda Y, Gonzales RA, Nelson
RS, Ding B. 2002. Potato spindle tuber vi-
roid strains of different pathogenicity in-
duces and suppresses expression of com-
mon and unique genes in infected tomato.
Mol. Plant-Microbe Interact. 15:990–
99
51. Kasschau KD, Xie Z, Allen E, Llave C,
Chapman EJ, et al. 2003. P1/HC-Pro, a
viral suppressor of RNA silencing, inter-
feres with Arabidopsis development and
miRNA function. Dev. Cell 4:205–17
52. Keese P, Symons RH. 1985. Domains in
viroids: evidence of intermolecular RNA
rearrangements and their contribution to
viroid evolution. Proc. Natl. Acad. Sci.
USA 82:4582–86
53. Khoury I, Singh RP, Boucher A, Coombs
DH. 1988. Concentration and distribution
of mild and severe strains of potato spin-
dle tuber viroid in cross-protected tomato
plants. Phytopathology 78:1331–36
54. Khvorova A, Lescoute A, Westhof E,
Jayasena SD. 2003. Sequence elements
outside the hammerhead ribozyme cat-
alytic core enable intracellular activity.
Nat. Struct. Biol. 10:708–12
55. Lecellier CH, Voinnet O. 2004. RNA si-
lencing: no mercy for viruses? Immunol.
Rev. 198:285–303
56. Lima MI, Fonseca MEN, Flores R, Ki-
tajima EW. 1994. Detection of avocado
sunblotch viroid in chloroplasts of avo-
cado leaves by in situ hybridization. Arch.
Virol. 138:385–90
57. Malfitano M, Di Serio F, Covelli L,
Ragozzino A, Hern´andez C, Flores R.
2003. Peach latent mosaic viroid variants
inducing peach calico contain a character-
istic insertion that is responsible for this
symptomatology. Virology 313:492–501
58. Maniataki E, Mart´ınez de Alba AE,
agesser R, Tabler M, Tsagris M. 2003.
Viroid RNA systemic spread may de-
pend on the interaction of a 71-nucleotide
bulged hairpin with the host protein VirP1.
RNA 9:346–54
59. Markarian N, Li HW, Ding SW, Semancik
JS. 2004. RNA silencing as related to vi-
roid induced symptom expression. Arch.
Virol. 149:397–406
60. Mart´ınez de Alba AE, Flores R,
Hern´andez C. 2002. Two chloroplastic vi-
roids induce the accumulation of the small
RNAs associated with post-transcrip-
tional gene silencing. J. Virol. 76:13094–
96
61. Mart´ınez de Alba AE, S¨agesser R, Tabler
M, Tsagris M. 2003. A bromodomain-
containing protein from tomato specif-
ically binds potato spindle tuber viroid
RNA in vitro and in vivo. J. Virol. 77:
9685–94
62. Matousek J, Orctova L, Steger G, Skopek
J, Moors M, et al. 2004. Analysis of ther-
mal stress-mediated PSTVd variation and
biolistic inoculation of progeny of viroid
“thermomutants” to tomato and Brassica
species. Virology 323:9–23
63. Mohamed NA, Thomas W. 1980. Viroid-
like properties of an RNA species associ-
ated with the sunblotch disease of avoca-
dos. J. Gen. Virol. 46:157–67
64. uhlbach HP, S¨anger HL. 1979. Viroid
replication is inhibited by α-amanitin. Na-
ture 278:185–88
65. Navarro B, Flores R. 1997. Chrysan-
themum chlorotic mottle viroid: unusual
structural properties of a subgroup of vi-
roids with hammerhead ribozymes. Proc.
Natl. Acad. Sci. USA 94:11262–67
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 137
66. Navarro JA, Flores R. 2000. Characteri-
zation of the initiation sites of both polar-
ity strands of a viroid RNA reveals a mo-
tif conserved in sequence and structure.
EMBO J. 19:2662–70
67. Navarro JA, Vera A, Flores R. 2000.
A chloroplastic RNA polymerase resis-
tant to tagetitoxin is involved in replica-
tion of avocado sunblotch viroid. Virology
268:218–25
68. Niblett CL, Dickson E, Fernow KH, Horst
RK, Zaitlin M. 1978. Cross-protection
among four viroids. Virology 91:198–203
69. Owens RA, Blackburn M, Ding B. 2001.
Possible involvement of the phloem lectin
in long-distance viroid movement. Mol.
Plant–Microbe Interact. 14:905–9
70. Owens RA, Steger G, Hu Y, Fels A, Ham-
mond RW, Riesner D. 1996. RNA struc-
tural features responsible for potato spin-
dle tuber viroid pathogenicity. Virology
222:144–58
71. Palacio-Bielsa A, Romero-Durban J,
Duran-Vila N. 2004. Characterization of
citrus HSVd isolates. Arch. Virol. 149:
537–52
72. Palukaitis P. 1987. Potato spindle tuber
viroid: investigation of the long-distance,
intra-plant transport route. Virology 158:
239–41
73. Papaefthimiou I, Hamilton AJ, Denti MA,
Baulcombe DC, Tsagris M, Tabler M.
2001. Replicating potato spindle tuber vi-
roid RNA is accompanied by short RNA
fragments that are characteristic of post-
transcriptional gene silencing. Nucleic
Acids Res. 29:2395–400
74. Park SW, Vepachedu R, Owens RA, Vi-
vanco JM. 2004. The N-glycosidase ac-
tivity of the ribosome-inactivating pro-
tein ME1 targets single-stranded regions
of nucleic acids independent of sequence
or structural motifs. J. Biol. Chem. 279:
34165–74
75. Pelchat M, Cˆot´eF,Perreault JP. 2001.
Study of the polymerization step of the
rolling circle replication of peach latent
mosaic viroid. Arch. Virol. 146:1753–63
76. Pelissier T, Wassenegger M. 2000. A
DNA target of 30 bp is sufficient for RNA-
directed DNA methylation. RNA 6:55–65
77. Pfeffer S, Zavolan M, Grasser FA, Chien
M, Russo JJ, et al. 2004. Identification
of virus-encoded microRNAs. Science
304:734–36
78. Prody GA, Bakos JT, Buzayan JM,
Schneider IR, Bruening G. 1986. Au-
tolytic processing of dimeric plant virus
satellite RNA. Science 231:1577–80
79. Qi Y, Ding B. 2002. Replication of potato
spindle tuber viroid in cultured cells of
tobacco and Nicotiana benthamiana: the
role of specific nucleotides in determin-
ing replication levels for host adaptation.
Virology 302:445–56
80. Qi Y, Ding B. 2003. Inhibition of cell
growth and shoot development by a spe-
cific nucleotide sequence in a noncoding
viroid RNA. Plant Cell 15:1360–74
81. Qi Y, Ding B. 2003. Differential sub-
nuclear localization of RNA strands of
opposite polarity derived from an au-
tonomously replicating viroid. Plant Cell
15:2566–77
82. Qi Y, Pelissier T, Itaya A, Hunt E,
Wassenegger M, Ding B. 2004. Direct role
of a viroid RNA motif in mediating direc-
tional RNA trafficking across a specific
cellular boundary. Plant Cell 16:1741–52
83. Reanwarakorn K, Semancik JS. 1998.
Regulation of pathogenicity in hop stunt
viroid-related group II citrus viroids. J.
Gen. Virol. 79:3163–71
84. Reid CE, Lazinski DW. 2000. A host-
specific function is required for liga-
tion of a wide variety of ribozyme-
processed RNAs. Proc. Natl. Acad. Sci.
USA 97:424–29
85. Rodriguez MJ, Randles JW. 1993. Co-
conut cadang-cadang viroid (CCCVd)
mutants associated with severe disease
vary in both the pathogenicity domain
and the central conserved region. Nucleic
Acids Res. 21:2771
86. anger HL. 1972. An infectious and repli-
cating RNA of low molecular weight: the
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
138 FLORES ET AL.
agent of exocortis disease of citrus. Adv.
Biosci. 8:103–16
87. anger HL, Klotz G, Riesner D, Gross HJ,
Kleinschmidt A. 1976. Viroids are single-
stranded covalently closed circular RNA
molecules existing as highly base-paired
rod-like structures. Proc. Natl. Acad. Sci.
USA 73:3852–56
88. Sano T, Matsuura Y. 2004. Accumula-
tion of short interfering RNAs character-
istic of RNA silencing precedes recovery
of tomato plants from severe symptoms
of potato spindle tuber viroid infection. J.
Gen. Plant Pathol. 70:50–53
89. Sano T, Candresse T, Hammond RW, Di-
ener TO, Owens RA. 1992. Identification
of multiple structural domains regulating
viroid pathogenicity. Proc. Natl. Acad.
Sci. USA 89:10104–8
90. Schiebel W, Pelissier T, Riedel L,
Thalmeir S, Schiebel R, et al. 1998.
Isolation of an RNA-directed RNA
polymerase-specific cDNA clone from
tomato. Plant Cell 10:2087–101
91. Schindler IM, M¨uhlbach HP. 1992. In-
volvement of nuclear DNA-dependent
RNA polymerases in potato spindle tuber
viroid replication: a reevaluation. Plant
Sci. 84:221–29
92. Schmitz A, Riesner D. 1998. Correlation
between bending of the VM region and
pathogenicity of different potato spindle
tuber viroid strains. RNA 4:1295–303
93. Schn¨olzer M, Haas B, Ramm K, Hof-
mann H, S¨anger HL. 1985. Correla-
tion between structure and pathogenic-
ity of potato spindle tuber viroid (PSTV).
EMBO J. 4:2181–90
94. Semancik JS, Conejero-Tomas V. 1987.
Viroid pathogenesis and expression of bi-
ological activity. In Viroids and Viroid-like
Pathogens, ed. JS Semancik, pp. 71–126.
Boca Raton: CRC Press
95. Semancik JS, Szychowski JA. 1994. Avo-
cado sunblotch disease: a persistent viroid
infection in which variants are associated
with differential symptoms. J. Gen. Virol.
75:1543–49
96. Semancik JS, Weathers LG. 1972. Exo-
cortis disease: evidence for a new species
of “infectious” low molecular weight
RNA in plants. Nat. New Biol. 237:242–
44
97. Singh RP, Clark MC. 1971. Infectious
low-molecular weight ribonucleic acid
from tomato. Biochem. Biophys. Res.
Commun. 44:1077–82
98. Solel Z, Mogilner N, Gafny R, Bar-Joseph
M. 1995. Induced tolerance to mal cecco
disease in etrog citron and Rangpur lime
by infection with citrus exocortis viroid.
Plant Dis. 79:60–62
99. Spiesmacher E, M¨uhlbach HP, Schn ¨olzer
M, Haas B, S¨anger HL. 1983. Oligomeric
forms of potato spindle tuber viroid
(PSTV) and of its complementary RNA
are present in nuclei isolated from viroid-
infected potato cells. Biosci. Rep. 3:767–
74
100. Tabler M, Tsagris M. 2004. Viroids: petite
RNA pathogens with distinguished tal-
ents. Trends Plant Sci. 9:339–48
101. Taylor JM. 2003. Replication of human
hepatitis delta virus: recent developments.
Trends Microbiol. 11:185–90
102. Tsagris M, Tabler M, M¨uhlbach HP,
anger HL. 1987. Linear oligomeric
potato spindle tuber viroid (PSTV) RNAs
are accurately processed in vitro to the
monomeric circular viroid proper when
incubated with a nuclear extract from
healthy potato cells. EMBO J. 6:2173–
83
103. Visvader JE, Symons RH. 1986. Replica-
tion of in vitro constructed viroid mutants:
location of the pathogenicity-modulating
domain in citrus exocortis viroid. EMBO
J. 13:2051–55
104. Vogt U, Pelissier T, Putz A, Razvi F, Fis-
cher R, Wassenegger M. 2004. Viroid-
induced RNA silencing of GFP-viroid fu-
sion transgenes does not induce extensive
spreading of methylation or transitive si-
lencing. Plant J. 1:107–18
105. Wang MB, Bian XY, Wu LM, Liu LX,
Smith NA, et al. 2004. On the role of RNA
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
26 Jul 2005 11:42 AR AR250-PY43-06.tex XMLPublishSM(2004/02/24) P1: KUV
VIROIDS AND VIROID-HOST INTERACTIONS 139
silencing in the pathogenicity and evolu-
tion of viroids and viral satellites. Proc.
Natl. Acad. Sci. USA 101:3275–80
106. Warrilow D, Symons RH. 1999. Citrus ex-
ocortis viroid RNA is associated with the
largest subunit of RNA polymerase II in
tomato in vivo. Arch. Virol. 144:2367–75
107. Wassenegger M, Heimes S, Riedel L,
anger HL. 1994. RNA-directed de novo
methylation of genomic sequences in
plants. Cell 76:567–76
108. Wassenegger M, Spieker RL, Thalmeir
S, Gast FU, Riedel L, S¨anger HL. 1996.
A single nucleotide substitution converts
potato spindle tuber viroid (PSTVd) from
a noninfectious to an infectious RNA for
Nicotiana tabacum.Virology 226:191–97
109. Woo Y-M, Itaya A, Owens RA, Tang L.
1999. Characterization of nuclear import
of potato spindle tuber viroid RNA in per-
meabilized protoplasts. Plant J. 17:627–
35
110. Zhao Y, Owens RA, Hammond RW.
2001. Use of a vector based on potato virus
Xinawhole plant assay to demonstrate
nuclear targeting of potato spindle tuber
viroid. J. Gen. Virol. 82:1491–97
111. Zhu Y, Green L, Woo Y-M, Owens R,
Ding B. 2001. Cellular basis of potato
spindle tuber viroid systemic movement.
Virology 279:69–77
112. Zhu Y, Qi Y, Xun Y, Owens R, Ding
B. 2002. Movement of potato spindle
tuber viroid reveals regulatory points
of phloem-mediated RNA traffic. Plant
Physiol. 130:138–46
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
VIROIDS AND VIROID-HOST INTERACTIONS C-1
Figure 1 Structural features of viroids. (a) Rod-like secondary structure of mem-
bers of the family Pospiviroidae. Domains C (central), P (pathogenic), V (variable),
and TLand TR(terminal left and right,respectively) are indicated, as well as the
CCR (central conserved region,here displayed for the genus Pospiviroid), TCR
(terminal conserved region,present in the genera Pospi- and Apscaviroid,and in the
two largest members of the genus Coleviroid) and TCH (terminal conserved hairpin,
present in the genera Hostu- and Cocadviroid). Arrows indicate flanking sequences
that together with the upper CCR strand form a hairpin, and the S-shaped line con-
nects the residues linked after UV irradiation as a consequence of forming part of the
loop E. (b) Quasi-rod-like and branched secondary structures of ASBVd and
PLMVd, respectively (family Avsunviroidae). Sequences conserved in most natural
hammerhead structures are shown within boxes with blue and white backgrounds for
() and () polarities, respectively. Broken oval in PLMVd denotes a kissing-loop
interaction. (Inset) PLMVd () hammerhead structure represented according to the
original scheme (left) and to X-ray crystallography data obtained with artificial ham-
merhead structures (right), in which a proposed tertiary interaction between loops 1
and 2 enhancing the catalytic activity is indicated. Nucleotides conserved in most
natural hammerhead structures are depicted as above. Arrows mark the self-cleavage
site, and continuous and broken lines denote Watson-Crick and noncanonical pairs,
respectively.
HI-RES-PY43-06-Flores.qxd 7/26/05 12:23 PM Page 1
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
C-2 FLORES ET AL.
Figure 2 Alterations in different organs accompanying infections by some repre-
sentative viroids. (a) Symptoms of PSTVd on potato tubers (left) and healthy con-
trols (right) (courtesy of T.O. Diener). (b) Symptoms of CEVd on a trifoliate orange
rootstock (courtesy of N. Duran-Vila and P. Moreno). (c) Symptoms of PBCVd on
pear A20 (courtesy of J.C. Desvignes). (d) Flower symptoms of CSVd on chrysan-
themum (bottom) and a healthy control (top) (courtesy of J.M. Bové). (e) Leaf
symptoms of CSVd on chrysanthemum (right) and a healthy control (left). ( f)Leaf
symptoms of CChMVd on chrysanthemum (right) and a healthy control (left).
(g)Leaf symptoms of CCCVd on coconut (left) and a healthy control (right).
(h) Extreme leaf chlorosis (peach calico) induced by PLMVd. (i) Internode shorten-
ing induced by HSVd on cucumber (right) and a healthy control (left). ( j)Symptoms
of HSVd on cucumber fruits (left) and a healthy control (right) (courtesy of H.L.
Sänger). (k) Fruit symptoms (dapple apple) induced by ASSVd (courtesy of J.C.
Desvignes). (l) Fruit symptoms of ASBVd on avocado (courtesy of P.R. Desjardins).
HI-RES-PY43-06-Flores.qxd 7/26/05 12:24 PM Page 2
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
VIROIDS AND VIROID-HOST INTERACTIONS C-3
Figure 3 Rolling-circle mechanism proposed for viroid replication. The asymme-
tric pathway, with one rolling circle, is followed by members of the family
Pospiviroidae and takes place in the nucleus. The symmetric pathway, with two
rolling circles, is followed by members of the family Avsunviroidae and takes place
in the chloroplast. White and yellow lines indicate plus () and minus () strands,
respectively, and cleavage sites are marked by arrowheads. Self-cleavage mediated
by hammerhead ribozymes (Rz) leads to linear monomeric RNAs with 5-hydroxyl
and 2-3-cyclic phosphodiester termini; cleavage catalyzed by a host protein (HP)
probably generates the same termini. Pol II refers to RNA polymerase II and NEP to
nuclear-encoded RNA polymerase. PD and NP are abbreviations for plasmodesma-
ta and nuclear pores, respectively.
HI-RES-PY43-06-Flores.qxd 7/26/05 12:24 PM Page 3
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
C-4 FLORES ET AL.
Figure 4 Hypothetical mechanisms of viroid pathogenesis. Symptoms could result
from direct interaction between the genomic viroid RNAs and a host factor (HF, pro-
tein or RNA), either in the organelle where the viroid replicates and accumulates or
in the cytoplasm during viroid movement. Alternatively, RNA silencing could medi-
ate symptom development. In the family Pospiviroidae viroid-specific siRNAs could
accumulate either in the nucleus (from the action of DCL1 on the genomic viroid
RNAs, or of DCL3 and/or DCL4 on the dsRNAs formed during replication), or in
the cytoplasm (from the action of DCL2 on the aberrant RNAs generated by a cyto-
plasmic RdRp, or on the genomic viroid RNAs, not shown). In the family
Avsunviroidae viroid-specific siRNAs should be generated only in the cytoplasm
because no chloroplastic DCL has been characterized. siRNAs would then load
RISC for degradation or translation repression of host mRNAs with complementary
sequences, or for direct methylation of host DNA. siRNAs could also target geno-
mic viroid RNAs and regulate their titer through a feed-back mechanism. The sizes
of the different molecules are not proportional. Both hypothetical mechanisms
should not be necessarily exclusive. PD, plasmodesmata; NP, nuclear pores.
HI-RES-PY43-06-Flores.qxd 7/26/05 12:24 PM Page 4
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
P1: KUV
July 14, 2005 11:17 Annual Reviews AR250-FM
Annual Review of Phytopathology
Volume 43, 2005
CONTENTS
FRONTISPIECE,Robert K. Webster xii
BEING AT THE RIGHT PLACE,AT THE RIGHT TIME,FOR THE RIGHT
REASONS—PLANT PATHOLOGY,Robert K. Webster 1
FRONTISPIECE,Kenneth Frank Baker
KENNETH FRANK BAKER—PIONEER LEADER IN PLANT PATHOLOGY,
R. James Cook 25
REPLICATION OF ALFAMO-AND ILARVIRUSES:ROLE OF THE COAT PROTEIN,
John F. Bol 39
RESISTANCE OF COTTON TOWARDS XANTHOMONAS CAMPESTRIS pv.
MALVACEARUM,E. Delannoy, B.R. Lyon, P. Marmey, A. Jalloul, J.F. Daniel,
J.L. Montillet, M. Essenberg, and M. Nicole 63
PLANT DISEASE:ATHREAT TO GLOBAL FOOD SECURITY,Richard N. Strange
and Peter R. Scott 83
VIROIDS AND VIROID-HOST INTERACTIONS,Ricardo Flores,
Carmen Hern´
andez, A. Emilio Mart´
ınez de Alba, Jos´
e-Antonio Dar`
os,
and Francesco Di Serio 117
PRINCIPLES OF PLANT HEALTH MANAGEMENT FOR ORNAMENTAL PLANTS,
Margery L. Daughtrey and D. Michael Benson 141
THE BIOLOGY OF PHYTOPHTHORA INFESTANS AT ITS CENTER OF ORIGIN,
Niklaus J. Gr¨
unwald and Wilbert G. Flier 171
PLANT PATHOLOGY AND RNAi: A BRIEF HISTORY,John A. Lindbo
and William G. Doughtery 191
CONTRASTING MECHANISMS OF DEFENSE AGAINST BIOTROPHIC AND
NECROTROPHIC PATHOGENS,Jane Glazebrook 205
LIPIDS,LIPASES,AND LIPID-MODIFYING ENZYMES IN PLANT DISEASE
RESISTANCE,Jyoti Shah 229
PATHOGEN TESTING AND CERTIFICATION OF VITIS AND PRUNUS SPECIES,
Adib Rowhani, Jerry K. Uyemoto, Deborah A. Golino,
and Giovanni P. Martelli 261
MECHANISMS OF FUNGAL SPECIATION,Linda M. Kohn 279
vii
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
P1: KUV
July 14, 2005 11:17 Annual Reviews AR250-FM
viii CONTENTS
PHYTOPHTHORA RAMORUM:INTEGRATIVE RESEARCH AND MANAGEMENT
OF AN EMERGING PATHOGEN IN CALIFORNIA AND OREGON FORESTS,
David M. Rizzo, Matteo Garbelotto, and Everett M. Hansen 309
COMMERCIALIZATION AND IMPLEMENTATION OF BIOCONTROL,D.R. Fravel 337
EXPLOITING CHINKS IN THE PLANTSARMOR:EVOLUTION AND EMERGENCE
OF GEMINIVIRUSES,Maria R. Rojas, Charles Hagen, William J. Lucas,
and Robert L. Gilbertson 361
MOLECULAR INTERACTIONS BETWEEN TOMATO AND THE LEAF MOLD
PATHOGEN CLADOSPORIUM FULVUM,Susana Rivas
and Colwyn M. Thomas 395
REGULATION OF SECONDARY METABOLISM IN FILAMENTOUS FUNGI,
Jae-Hyuk Yu and Nancy Keller 437
TOSPOVIRUS-THRIPS INTERACTIONS,Anna E. Whitfield, Diane E. Ullman,
and Thomas L. German 459
HEMIPTERANS AS PLANT PATHOGENS,Isgouhi Kaloshian
and Linda L. Walling 491
RNA SILENCING IN PRODUCTIVE VIRUS INFECTIONS,Robin MacDiarmid 523
SIGNAL CROSSTALK AND INDUCED RESISTANCE:STRADDLING THE LINE
BETWEEN COST AND BENEFIT,Richard M. Bostock 545
GENETICS OF PLANT VIRUS RESISTANCE,Byoung-Cheorl Kang, Inhwa Yeam,
and Molly M. Jahn 581
BIOLOGY OF PLANT RHABDOVIRUSES,Andrew O. Jackson, Ralf G. Dietzgen,
Michael M. Goodin, Jennifer N. Bragg, and Min Deng 623
INDEX
Subject Index 661
ERRATA
An online log of corrections to Annual Review of Phytopathology chapters
may be found at http://phyto.annualreviews.org/
Annu. Rev. Phytopathol. 2005.43:117-139. Downloaded from www.annualreviews.org
by INRA Institut National de la Recherche Agronomique on 05/16/13. For personal use only.
... They are long non-coding RNA molecules composed of 246-434 nucleotides (nts) and have a singlestranded, covalently closed circular genome [1]. Viroids infect plant cells through mechanical means, replicate and move within the host, and can cause serious diseases in vegetables, flowers, and fruit trees [2][3][4]. Citrus is a perennial woody plant in the Rutaceae family that is infected by various viroids. So far, citrus is a natural host of at least eight viroid species, including Citrus exocortis viroid (CEVd), Citrus bent leaf viroid (CBLVd), Hop stunt viroid (HSVd), Citrus dwarfing viroid (CDVd), Citrus bark cracking viroid (CBCVd), Citrus viroid V (CVd-V), Citrus viroid VI (CVd-VI), and Citrus viroid VII (CVd-VII). ...
... The titer of viroids can be co-regulated by DCLs and the RISC. DCLs process genomic RNA and the replication intermediates of viroids, and the resulting siRNAs are assembled into RISC to target and degrade the viroid RNA [2]. A similar mechanism can explain the antagonism among citrus viroids. ...
Article
Full-text available
Citrus is the natural host of at least eight viroid species, providing a natural platform for studying interactions among viroids. The latter manifests as antagonistic or synergistic phenomena. The antagonistic effect among citrus viroids intuitively leads to reduced symptoms caused by citrus viroids, while the synergistic effect leads to an increase in symptom severity. The interaction phenomenon is complex and interesting, and a deep understanding of the underlying mechanisms induced during this viroid interaction is of great significance for the prevention and control of viroid diseases. This paper summarizes the research progress of citrus viroids in recent years, focusing on the interaction phenomenon and analyzing their interaction mechanisms. It points out the core role of the host RNA silencing mechanism and viroid-derived siRNA (vd-siRNA), and provides suggestions for future research directions.
... Viroids are fascinating biological entities characterized by their extremely simple genomes, short (between 200 and 400-nt) circular non-coding RNAs, that are only pathogenic to plants (1,2). Classified as sub-viral pathogens, viroids can be subdivided into two families according to their ability to replicate in the nucleus (Pospiviroidae) or the chloroplast (Avsunviroidae) (3,4). ...
... Due to their remarkably simple genomic organization, viroids must interact with the molecular machinery of the plant cell to fulfill every aspect of their life cycle. As a consequence of this extremely close interaction with their host, viroids induce developmental defects that are identified as symptomatology and are very similar to symptoms caused by viruses (1,2). ...
Preprint
Full-text available
Viroids are pathogenic non-coding RNAs that completely rely on their host molecular machinery to accomplish their life cycle. Several interactions between viroids and their host molecular machinery have been identified, including an interference with epigenetic mechanisms such as DNA methylation. Despite this, whether viroids influence changes in other epigenetic marks such as histone modifications remained unknown. Epigenetic regulation is particularly important during pathogenesis processes because it might be a key regulator of the dynamism of the defense response. Here we have analyzed the changes taking place in Cucumis sativus facultative and constitutive heterochromatin during hop stunt viroid (HSVd) infection using chromatin immunoprecipitation (ChIP) of the two main heterochromatic marks: H3K9me2 and H3K27me3. We find that HSVd infection is associated with changes in both H3K27me3 and H3K9me2, with a tendency to decrease the levels of repressive epigenetic marks through infection progression. These epigenetic changes are connected to the transcriptional regulation of their expected targets, genes and transposable elements. Indeed, several genes related to the defense response are targets of both epigenetic marks. Our results highlight another host regulatory mechanism affected by viroid infection, providing further information about the complexity of the multiple layers of interactions between pathogens/viroids and hosts/plants.
... In this study, potato spindle tuber viroid (PSTVd) was chosen as a model to achieve a more comprehensive understanding of the relationship between RNA secondary structure and function. The RNA genome of PSTVd comprises 359 nucleotides and adopts a rod-like structure characterized by 27 loops and 26 stems or helices [12][13][14]. After undergoing replication within the nucleus, facilitated by the host RNA polymerase II, PSTVd traverses between cells via plasmodesmata and can travel over long distances within the phloem. ...
... However, only twelve positions were chosen to effectively represent various regions of the genome and different Figure 1A). The secondary structure of the PSTVd intermediate strain resembles a rod-like shape, the five common domains found in members of the Pospiviroidae family were presented ( Figure 1A) [12][13][14][15]. The starting sites of some of the 12 forms are mapped to the two conserved regions. ...
Article
Full-text available
The function of RNAs is determined by their structure. However, studying the relationship between RNA structure and function often requires altering RNA sequences to modify the structures, which leads to the neglect of the importance of RNA sequences themselves. In our research, we utilized potato spindle tuber viroid (PSTVd), a circular-form non-coding infectious RNA, as a model with which to investigate the role of a specific rod-like structure in RNA function. By generating linear RNA transcripts with different start sites, we established 12 PSTVd forms with different secondary structures while maintaining the same sequence. The RNA secondary structures were predicted using the mfold tool and validated through native PAGE gel electrophoresis after in vitro RNA folding. Analysis using plant infection assays revealed that the formation of a correct rod-like structure is crucial for the successful infection of PSTVd. Interestingly, the inability of PSTVd forms with non-rod-like structures to infect plants could be partially compensated by increasing the amount of linear viroid RNA transcripts, suggesting the existence of additional RNA secondary structures, such as the correct rod-like structure, alongside the dominant structure in the RNA inoculum of these forms. Our study demonstrates the critical role of RNA secondary structures in determining the function of infectious RNAs.
... Viroids are the smallest pathogens in molecular weight comprised of circular single-stranded RNA without a protein coat; they are currently the smallest known pathogens. Pospiviroidae and Avsunviroidae have been classified as viroids (Adkar-Purushothama & Perreault, 2020;Ding, 2009;Flores et al., 2005). Potato spindle tuber viroid (PSTVd) is the first identified individual of the Pospiviroidae (Apostolova et al., 2020). ...
Article
Full-text available
Tomato is a popular vegetable worldwide; its production is highly threatened by infection with the potato spindle tuber viroid (PSTVd). We obtained the full‐length genome sequence of previously conserved PSTVd and inoculated it on four genotypes of semi‐cultivated tomatoes selected from a local tomato germplasm resource. SC‐5, which is a PSTVd‐resistant genotype, and SC‐96, which is a PSTVd‐sensitive genotype, were identified by detecting the fruit yield, plant growth, biomass accumulation, physiological indices, and PSTVd genome titer after PSTVd inoculation. A non‐target metabolomics study was conducted on PSTVd‐infected and control SC‐5 to identify potential anti‐PSTVd metabolites. The platform of liquid chromatography‐mass spectrometry detected 158 or 123 differential regulated metabolites in modes of positive ion or negative ion. Principal component analysis revealed a clear separation of the global metabolite profile between PSTVd‐infected leaves and control regardless of the detection mode. The potential anti‐PSTVd compounds, xanthohumol, oxalicine B, indole‐3‐carbinol, and rosmarinic acid were significantly upregulated in positive ion mode, whereas echinocystic acid, chlorogenic acid, and 5‐acetylsalicylic acid were upregulated in negative ion mode. Xanthohumol and echinocystic acid were detected as the most upregulated metabolites and were exogenously applied on PSTVd‐diseased SC‐96 seedlings. Both xanthohumol and echinocystic acid had instant and long‐term inhibition effect on PSTVd titer. The highest reduction of disease symptom was induced by 2.6 mg/L of xanthohumol and 2.0 mg/L of echinocystic acid after 10 days of leaf spraying, respectively. A superior effect was seen on echinocystic acid than on xanthohumol. Our study provides a statistical basis for breeding anti‐viroid tomato genotypes and creating plant‐originating chemical preparations to prevent viroid disease.
Article
The symptoms caused by viroids differ, ranging from asymptomatic to mild-or-severe symptoms. Pepper plant symptoms caused by the Pepper chat fruit viroid (PCFVd) are mild compared to those affecting tomato plants; however, there is not much more known of the symptomatology on pepper plants. Symptoms that could be used for disease virulence assessment in pepper plants were elucidated from 31 commercial pepper cultivars belonging to Capsicum annuum and C. frutescens that had been purchased from agricultural shops. The plants were mechanically sap-inoculated at the seedling growth stage and observed weekly for symptom development, with disease virulence evaluations performed at 4, 8, and 12 weeks post inoculation. Infection of all plants was verified based on reverse transcription-polymerase chain reaction, along with visual evidence of growth reduction, including leaf rugosity and leaf size reduction, a narrow canopy of the blocky and elongated fruit shapes for C. annuum and explicit apical stunting with small apical leaves of the elongated fruit type for C. frutescens. The disease virulence assessment was designed based on these symptoms to produce a score with 0–10 disease virulence levels (DVLs). The results showed that the pepper cultivars displayed responses to PCFVd with DVL scores of 1.00–8.00, with no PCFVd transmission being recorded from seeds to seedlings for the 3 test cultivars. This finding indicated that the genetic resources of pepper cultivars against PCFVd were as low as 1.00 DVL. However, the low DVL pepper cultivars could provide an inoculum source to other susceptible plants via mechanical transmission.
Article
Full-text available
Citrus hosts various phytopathogens that have impacted productivity, including viroids. Missing data on the status of viroids in citrus in Palestine were not reported. This study was aimed to detect any of Citrus exocortis viroid (CEVd), Citrus viroid-III (CVd-III), and Citrus viroid-IV (CVd-IV) in the Palestinian National Agricultural Research Center (NARC) germplasm collection Field inspections found symptoms such as leaf epinasty; vein discoloration, and bark cracking on various citrus varieties. RT-PCR revealed a significant prevalence of CVd-IV; CEVd and CVd-III (47%, 31%, and 22%; respectively). CVd-III variants with 91.3% nucleic acid sequence homology have been reported. The sequence of each viroid were deposited in GenBank as (OP925746 for CEVd, OP902248 and OP902249 for CVd-III-PS-1 and -PS-2 isolates, and OP902247 for CVd-IV). This was the first to report three of citrus viroids in Palestine, appealing to apply of phytosanitary measures to disseminate healthy propagating materials free from viroids.
Article
Full-text available
The increased cultivation of Cannabis sativa L. in North America, represented by high Δ9-tetrahydrocannabinol-containing (high-THC) cannabis genotypes and low-THC-containing hemp genotypes, has been impacted by an increasing number of plant pathogens. These include fungi which destroy roots, stems, and leaves, in some cases causing a build-up of populations and mycotoxins in the inflorescences that can negatively impact quality. Viroids and viruses have also increased in prevalence and severity and can reduce plant growth and product quality. Rapid diagnosis of the occurrence and spread of these pathogens is critical. Techniques in the area of molecular diagnostics have been applied to study these pathogens in both cannabis and hemp. These include polymerase chain reaction (PCR)-based technologies, including RT-PCR, multiplex RT-PCR, RT-qPCR, and ddPCR, as well as whole-genome sequencing (NGS) and bioinformatics. In this study, examples of how these technologies have enhanced the rapidity and sensitivity of pathogen diagnosis on cannabis and hemp will be illustrated. These molecular tools have also enabled studies on the diversity and origins of specific pathogens, specifically viruses and viroids, and these will be illustrated. Comparative studies on the genomics and metabolomics of healthy and diseased plants are urgently needed to provide insight into their impact on the quality and composition of cannabis and hemp-derived products. Management of these pathogens will require monitoring of their spread and survival using the appropriate technologies to allow accurate detection, followed by appropriate implementation of disease control measures.
Article
Sequence analysis by primer‐extension at the level of their cDNA showed that the RNA genomes of various field isolates of potato spindle tuber viroid (PSTV) of different virulence differ from each other only in a few nucleotides in two distinct regions of the rod‐shaped molecule. Despite insertions and deletions the chain length of 359 nucleotides is strictly conserved in all the isolates studied. Thermodynamic calculations revealed that due to the observed sequence differences the region located at the left hand part of the rod‐like secondary structure of the PSTV molecule, denoted ‘virulence modulating (VM) region’, becomes increasingly unstable with the increasing virulence of the corresponding isolate. Based on these data we propose in molecular terms a model for the mechanism of viroid pathogenicity. It implies that the nucleotides of the VM region specify and modulate the binding‐ and hence the competition‐potential of the PSTV RNA molecule for a still unknown host factor(s) and thus determine the virulence of PSTV.
Article
The transmissible agent of the exocortis disease of citrus (ExC) exhibits properties of an unusual free infectious RNA. It can be isolated from its herbaceous host Gynura aurantiaca DC by any method designed to extract and to concentrate undegraded host or viral RNA. The infectivity of ExC is found in nuclei and in association with the chromatin of the host cell. The agent of ExC is insensitive to different organic solvents, to proteolytic enzymes, and to DNase, but it is inactivated by RNase and by formaldehyde. In rate sedimentation experiments, it sedimented with about 6 to 8 s. Electrophoretic analysis in appropriate polyacrylamide gels revealed that infectivity of ExC is consistently associated with a monodisperse fraction of RNA with a molecular weight of ca. 50–60.000 daltons which would comprise molecules with a chain length of ca. 150–200 nucleotides.
Chapter
The sunblotch disease of avocado has been known for over 50 years. It was first described as a physiological (Coit, 1928) or genetic (Home, 1929) disorder. Shortly thereafter, Horne and Parker (1931) transmitted the causal agent from diseased scions to healthy rootstocks. By 1941, graft transmissibility of the agent was well established (Horne et al., 1941). Hence, the causal agent was considered to be a virus, and like some of the other plant viroids was studied as such for many years (Whitsell, 1952; Wallace, 1958; Wallace and Drake, 1962; Desjardins et al., 1980).
Article
A 3600-bp RNA-directed RNA polymerase (RdRP)–specific cDNA comprising an open reading frame (ORF) of 1114 amino acids was isolated from tomato. The putative protein encoded by this ORF does not share homology with any characterized proteins. Antibodies that were raised against synthetic peptides whose sequences have been deduced from the ORF were shown to specifically detect the 127-kD tomato RdRP protein. The immunoresponse to the antibodies correlated with the enzymatic activity profile of the RdRP after chromatography on Q-, poly(A)–, and poly(U)–Sepharose, hydroxyapatite, and Sephadex G-200 columns. DNA gel blot analysis revealed a single copy of the RdRP gene in tomato. RdRP homologs from petunia, Arabidopsis, tobacco, and wheat were identified by using polymerase chain reaction. A sequence comparison indicated that sequences homologous to RdRP are also present in the yeast Schizosaccharomyces pombe and in the nematode Caenorhabditis elegans. The previously described induction of RdRP activity upon viroid infection is shown to be correlated with an increased steady state level of the corresponding mRNA. The possible involvement of this heretofore functionally elusive plant RNA polymerase in homology-dependent gene silencing is discussed.
Article
Etrog citron and Rangpur lime plants were each inoculated with one of four isolates of the citrus exocortis viroid (CEVd) and incubated for 1 yr, during which time the viroid spread systemically. CEVd caused severe leaf curling and short internodes in Etrog citron but not in Rangpur lime. After leaf inoculation with Phoma tracheiphila, typical mal secco disease symptoms developed on leaves of both cultivars, regardless of prior infection by CEVd, but growth of the mycelium from the infected leaves into the branches was greatly affected by CEVd infection. In Etrog citron, P. tracheiphila was isolated from only 9.1% of the branches sampled from CEVd-infected plants, compared with 100% from CEVd-free plants. In Rangpur lime, the incidence of branches containing P. tracheiphila varied in CEVd-infected plants from 16.7 to 50% among CEVd isolates, compared with 70% in the viroid-free controls. This is believed to be the first report of systemic resistance induced in a woody plant by prior inoculation with a viroid
Article
Earlier results had indicated that the spindle tuber disease of potato is incited by free RNA, and that neither conventional virions nor proteins that could be construed as viral coat proteins are synthesized in infected plants. By a combination of density-gradient centrifugation and polyacrylamide gel electrophoresis, using internal marker RNAs, it is now shown that the infectious RNA occurs in the form of several species with molecular weights ranging from 2.5 × 104 to 1.1 × 105 daltons. No evidence for the presence in uninoculated plants of a latent helper virus was found. Thus, potato spindle tuber “virus” RNA, which is too small to contain the genetic information necessary for self-replication, must rely for its replication mainly on biosynthetic systems already operative in the uninoculated plant. Several possible mechanisms are discussed. The term “viroid” is proposed to designate potato spindle tuber “virus” RNA and other RNAs with similar properties.
Article
In the quest for plant regulatory sequences capable of driving nematode-triggered effector gene expression in feeding structures, we show that promoter tagging is a valuable tool, A large collection of transgenic Arabidopsis plants was generated, They were transformed with a beta-glucuronidase gene functioning as a promoter tag, Three T-DNA constructs, pGV1047, p Delta gusBin19, and pMOG553, were used. Early responses to nematode invasion were of primary interest. Six lines exhibiting beta-glucuronidase activity in syncytia induced by the beet cyst nematode were studied. Reporter gene activation was also identified in galls induced by root knot and ectoparasitic nematodes. Time-course studies revealed that all six tags were differentially activated during the development of the feeding structure. T-DNA-flanking regions responsible for the observed responses after nematode infection were isolated and characterized for promoter activity.