throbber
insight review articles
`
`Gene silencing as an adaptive
`defence against viruses
`
`Peter M. Waterhouse*, Ming-Bo Wang* & Tony Lough†
`
`*CSIRO Plant Industry, Canberra, ACT 2601, Australia (e-mail: peterw@pi.csiro.au)
`†HortResearch, Tennent Drive, Palmerston North, New Zealand (Present address: Genesis Research and Development Corporation, PO Box 50,
`Auckland, New Zealand)
`
`Gene silencing was perceived initially as an unpredictable and inconvenient side effect of introducing
`transgenes into plants. It now seems that it is the consequence of accidentally triggering the plant’s adaptive
`defence mechanism against viruses and transposable elements. This recently discovered mechanism,
`although mechanistically different, has a number of parallels with the immune system of mammals.
`
`Biology students are taught that the concept of
`
`vaccination came from Edward Jenner’s
`discovery that milkmaids and dairymen
`infected with the mild cowpox virus were
`protected against smallpox. It is less widely
`appreciated that plants can also be protected from a severe
`virus by prior infection with a mild strain of a closely
`related virus. This cross protection in plants was
`recognized as early as the 1920s, but its mechanism has
`been a mystery — plants do not possess an antibody-
`based immune system analogous to that found in animals.
`This was probably the first observation of a plant’s
`intrinsic defence mechanism against viruses which, 75
`years later, is just beginning to be understood. In the past
`decade, there has been considerable research into
`transgene-mediated virus resistance, co-suppression,
`virus-induced
`gene
`silencing
`(VIGS),
`antisense
`suppression and transcriptional gene silencing (TGS) in
`plants. There has also been intense research into RNA
`interference in Drosophila and nematodes, and quelling in
`fungi. These seemingly disparate endeavours have
`produced pieces of a jigsaw puzzle which, when put
`together, begin to reveal the existence and characteristics
`of a natural defence system in plants against viruses and
`transposable DNA elements. Many of the details
`and ramifications have yet to be determined, but the
`current picture is that of a wonderfully elegant system that
`can generically
`recognize
`invading viruses and
`transposable elements (TEs) and marshal the plant’s
`defences against them.
`
`Plant viruses and transposable DNA elements
`There are currently 72 different defined genera of plant
`viruses1, containing over 500 species, and there is scarcely a
`plant species — mono- or dicotyledon — that is not host to
`at least one virus. Plant viruses have a whole array of
`different particle morphologies, host ranges, vectors (for
`example, insects, nematodes, fungi, pollen, seeds or
`humans), genome organizations and gene expression
`strategies. They cause symptoms which at their least severe
`are unnoticeable, but range upwards through ringspots or
`mosaic leaf patterns, to widespread necrosis. The genomes
`of some plant viruses are encoded using single-stranded
`(ss) or double-stranded (ds) DNA; others have dsRNA
`genomes. However, over 90% of plant viruses have ssRNA
`genomes
`that are replicated by a virus-encoded
`RNA-dependent RNA polymerase (RDRP). Plants defend
`
`themselves by exploiting this requirement of most plant
`viruses to replicate using a double-stranded replicative
`intermediate.
`TEs are DNA sequences that have the capacity to move
`from place to place within a genome. They have been
`divided into two classes. Class I TEs are retroelements that
`amplify their copy number through reverse transcription of
`an RNA intermediate. They are particularly abundant in
`eukaryotes, and in plants comprise the greatest mass of TEs
`(in maize, this class of TE makes up over 70% of the nuclear
`DNA). Of the four types of retroelements in plants, the
`main class contains retrotransposons with direct long
`terminal repeats (LTRs). Class II TEs occur in all organisms,
`particularly prokaryotes; they have terminal inverted
`repeats (TIRs) ranging in size from 11 to several hundred
`base pairs. Within a class II TE family, one or more elements
`encode a transposase that has the potential to interact with
`TIRs to excise the elements and integrate them into other
`regions of the genome (for recent reviews of plant TEs, see
`refs 2, 3). Both classes of TEs can move around plant
`genomes, altering the function and structure of genes, and
`so accelerating genomic evolution. However, they are also
`parasitic mutagenic agents that have the potential to
`lacerate a genome4. To ensure survival, a plant needs to keep
`TE activity in check.
`
`Targeted RNA degradation
`Although not recognized at the time, evidence of a plant’s
`intrinsic defence mechanism to counter viruses and
`transposons came from the initially mystifying results of co-
`suppression and transgene-mediated virus resistance.
`Transformation with antisense gene constructs has been
`used in plant research since 1987 (ref. 5). From previous
`work on natural antisense in bacteria6, it was thought that
`hybridization of antisense RNA to the target messenger
`RNA interfered with its transport or translation7. So it was
`surprising to find subsequently that transformation of
`plants with transgene constructs encoding sense mRNA
`homologous to endogenous genes could also suppress the
`activities of these genes8–10. It was similarly perplexing when
`plants
`transformed with virus-derived
`transgenes,
`designed to provide protection through a protein-mediated
`mechanism11, gave protection against viruses even
`when little or no transgene protein (transprotein) was
`produced12 (Fig. 1).
`Further analysis of the co-suppressed and virus-
`resistant transprotein-free plants revealed that in both cases
`
`834
`
`© 2001 Macmillan Magazines Ltd
`
`NATURE | VOL 411 | 14 JUNE 2001 | www.nature.com
`
`Benitec - Exhibit 1013 - page 1
`
`

`
`insight review articles
`
`consistent with the phenomenon of VIGS; here, plant viruses that
`contain sequences homologous to nuclear-expressed genes act to
`induce silencing of the targeted genes26.
`It therefore seems likely that one of the roles of the dsRNA-
`induced RNA degradation system of plants is to protect them against
`virus infection. This is a way to generically detect the replication of an
`invading ssRNA virus and destroy it, by specifically degrading both
`replicating and translatable forms of its genome.
`
`Nuclease specificity and location
`Specific fragments from mRNAs or viral genomes have been
`identified in gene-silenced or virus-resistant tissues, indicating
`that the targeted RNA degradation starts with endonucleolytic
`cleavage at one or more sites and is followed by exonucleolytic
`degradation14,27,28. Further investigation has found that sense and
`antisense ~25-nucleotide RNAs, with homology to the target RNA,
`are found consistently in plants showing co-suppression, antisense
`suppression, VIGS and virus resistance, but not in the appropriate
`control plants29–33. This is a critical finding. It supplies further
`evidence that these different forms of silencing are all acting by the
`same mechanism; from here on we refer to them generically as post-
`transcriptional gene silencing (PTGS). It also provides a strong link
`between PTGS and a phenomenon called RNA interference (RNAi),
`which is the targeted inhibition of gene activity by introduction,
`usually by injection, of dsRNA into a number of lower eukaryotes,
`including nematodes and Drosophila34,35.
`Looking at the biochemistry of the process of RNAi provides a
`good indication of what is probably happening in plants (Fig. 2).
`The processes of RNAi have been examined in Drosophila embryos,
`and embryo extracts, using radiolabelled dsRNA and target
`ssRNA36–40. Target ssRNA is not significantly degraded when sense
`or antisense RNAs are also introduced. However, the target RNA is
`degraded within minutes of adding homologous dsRNA. The degra-
`dation rapidly produces short sense and antisense ~21-nucleotide
`RNAs from both the dsRNA and the target ssRNA as a two-step
`process39,40. The dsRNA is degraded in ~21-nucleotide steps from
`both ends by an enzyme called Dicer-1 (CG64792, DCR1). The
`cleavage process, which is similar to that of Escherichia coli RNase III,
`produces ~21-nucleotide dsRNA fragments with 38 overhangs of
`2–3 nucleotides, and 58-phosphate and 38-hydoxyl termini40. Each
`fragment is associated with, and cleaved by, a separate Dicer-
`containing complex. The current model for the second step of the
`degradation
`is that the Dicer-containing, small
`interfering
`ribonucleoprotein (siRNP) complex alters in such a way that the
`strands of short dsRNA become unpaired and guide the complex to
`complementary target RNAs. This probably requires recruitment of
`
`Figure 1 Potato plants
`challenged with potato virus Y.
`The three plants on the right
`are non-transgenic and are
`susceptible to the virus. The
`three plants on the left contain
`an untranslatable virus-
`derived transgene yet are
`immune to the virus18.
`
`the transgenes were being highly transcribed in the nucleus, but the
`steady-state levels of their mRNAs in the cytoplasm were very low.
`This led to the proposal12,13 that the transgene mRNA was somehow
`perceived by the cell as unwanted and induced sequence-specific
`degradation, by a targeted nuclease, of itself and other homologous
`or complementary RNA sequences in the cytoplasm. Thus, in the co-
`suppressed and virus-resistant lines, not only the transgene mRNAs
`but also the mRNA from the homologous endogenous gene and the
`invading virus RNA (with homology to the transgene) were being
`degraded. The concept of transgene RNA-directed RNA degradation
`was supported by the results of an experiment in which plants, with a
`co-suppressed b-glucuronidase (GUS) reporter gene, were inoculat-
`ed with a wild-type plant virus or the same virus engineered to
`contain GUS sequences in its genome. The plants were susceptible to
`the wild-type virus, but resistant to the virus containing the
`GUS-encoding sequence. The virus has an RNA genome and
`replicates exclusively in the cytoplasm, so the simple explanation is
`that the GUS sequence within the virus genome was specifically
`degraded in the cytoplasm by the same mechanism that was causing
`co-suppression of the nuclear-expressed genes14.
`This conclusion raised a number of questions. How do the
`nucleases in the cell know which RNAs to degrade and which to leave
`alone, or more specifically, how do they distinguish transgene RNA
`from endogenous gene mRNA? Why does this not happen to the
`mRNAs from all transgenes? And why would a plant want to
`specifically degrade these RNAs? A critical observation was that in
`both the co-suppression and virus-transgene transformation experi-
`ments, only a proportion of the initial transformants showed co-
`suppression or virus resistance, and these plants generally contained
`multiple, methylated copies of the transgenes.
`
`How is the degradation system triggered?
`Several theories have been advanced to explain how this sequence-
`specific degradation system might be activated. It was proposed
`initially that the high copy number of the transgenes produced
`excessively high levels of transgene mRNA and that this level induced
`the degradation system12,13. Other researchers suggested that the
`methylation of the transgenes made them produce aberrant (for
`example, prematurely terminated) RNA and that this aberrance
`induced the system14–17. A compelling proposal was that the system is
`induced and directed by dsRNA and that multiple transgenes
`favoured the likelihood of their integration as inverted repeats which,
`by transcriptional readthrough from one transgene into the other,
`would produce duplex-forming, self-complementary RNA18. This
`was supported by the demonstration that transgenes deliberately
`designed to produce self-complementary (hairpin or hp) RNA or
`dsRNA were highly efficient at inducing targeted virus resistance and
`gene silencing18–20. Furthermore, an investigation of simple co-
`suppression and antisense constructs found a perfect correlation
`between the integration of these constructs as inverted repeats and
`the induction of silencing19, and analysis of similar loci detected the
`presence of hpRNAs transcribed from them21.
`Why would the plant want to degrade dsRNA and ssRNAs of simi-
`lar sequence? Healthy plants do not contain dsRNA or extensively
`self-complementary ssRNA. In fact, for many years, plant virologists
`have used the presence of dsRNA in plant extracts to diagnose viral
`infection22. This seems to be the key. Most plant viruses have ssRNA
`genomes and replicate in the cytoplasm using their own RDRP to
`produce both sense and antisense (termed plus-strand and minus-
`strand) RNA. Evidence from research on the RNA bacteriophage
`Qb23 suggests that the plus and minus strands of a ssRNA virus form
`full-length dsRNA only as an artefact of extraction24. However, when
`the mammalian 28,58-oligoadenylate system (which is activated
`specifically by dsRNA) was transformed into plants, it was activated
`by infection with either of the two ssRNA plant viruses tested25.
`Therefore, replication of a ssRNA plant virus can produce sufficient
`stretches of dsRNA to be recognized as such within the plant. This is
`
`NATURE | VOL 411 | 14 JUNE 2001 | www.nature.com
`
`© 2001 Macmillan Magazines Ltd
`
`835
`
`Benitec - Exhibit 1013 - page 2
`
`

`
`Table 1 Genetic components of post-transcriptional silencing pathways
`Gene function Plants*
`Worms*
`Flies*
`Fungi*
`Algae*
`TE act.† PAZ‡
`RDRP
`SGS2
`EGO1
`QDE1
`SDE1§
`AGO1
`
`RDE1
`
`QDE2
`
`No
`
`Yes
`
`insight review articles
`
`a
`
`b
`
`c
`
`d
`
`e
`
`Translation
`initiation factor
`RecQ, DEAH
`or Upf1p
`helicase
`RNaseIII&
`helicase¶
`Chromatin
`remodelling
`Methyl
`transferase
`?
`?
`?
`
`SDE3
`
`MUT7||
`SMG-2
`
`QDE3
`
`MUT6|| Yes||
`
`CAF1
`
`K12H4.8 DCR1
`
`Yes
`
`DDM1
`
`MET1
`
`SGS3
`
`Yes
`
`No
`Yes
`
`RDE 4
`RDE2&3
`MUT2
`No
`No
`No
`Yes?
`Methylation
`*Plants, Arabidopsis thaliana; worms, Caenorhabditis elegans; flies, Drosophila melanogaster;
`fungi, Neurospora crassa; algae, Chlamydomonas reinhardtii.
`†TE act., activation of transposon activity in organisms mutant for this gene function.
`‡PAZ, presence of the PAZ domain identified by Cerruti et al.54.
`§SGS2 and SDE1 are different descriptions of the same gene.
`||The effect on transposon activation was assessed only for the MUT7 and MUT6 helicases.
`MUT7 also has an RNAse D motif.
`¶CAF1 and K12H4.8 have been identified by sequence homology to DCR1 but they have not
`been shown formally to be involved in PTGS.
`
`Figure 2 Proposed mechanism, based on RNAi, for dsRNA-directed ssRNA cleavage in
`PTGS. Introduced dsRNA (a) attracts Dicer-1-like proteins to its termini (b). The
`heterodimer complex at each end cleaves a 21-nucleotide dsRNA fragment (c), and the
`exposed ends of the shortened dsRNA each attract a new Dicer complex, which cleaves
`a further 21-nucleotide dsRNA fragment. This progressive exonuclease-like shortening
`continues until the dsRNA is completely cleaved. Dicers, loaded with dsRNA, acquire
`further components (blue ellipse), melt their dsRNA fragments and use one strand to
`hybridize to homologous ssRNA and cleave it in the middle of the 21-nucleotide guide-
`recognized sequence. One half of the dimer (shown as gold) directs hybridization and
`endonuclease cleavage giving it sense specificity. Thus Dicer complexes loaded from
`one end of a dsRNA will cleave sense-strand mRNA (d), whereas complexes loaded
`from the other end will cleave mRNA of the opposite sense (e).
`
`additional proteins to the complex39. Once hybridized to a target
`RNA, the complex cleaves it at a position approximately in the
`middle of the recognized 21-nucleotide sequence. The whole two-
`step process results in dsRNA being cleaved with a ~21-nucleotide
`(two helical turns36) periodicity from their termini, and the
`appropriate target RNAs being cleaved with the same periodicity but
`with a frame shift of ~10 nucleotides.
`The second step of the degradation probably takes place exclu-
`sively in the cytoplasm, as silencing does not reduce the full-length
`transcript levels in the nucleus27. However, the first step could occur
`in both the cytoplasm and the nucleus. Many mRNA degradation
`mechanisms involve the association of RNA with ribosomes, so it
`might be assumed that this would be the site of siRNP-mediated
`degradation. But several studies using protein-synthesis inhibitors
`have shown that neither ongoing translation nor association of the
`target RNAs with the ribosome are required for this degradation15,41.
`Furthermore, for each ribosome to be associated with enough siRNP
`complexes to ensure effective degradation of target RNA, the siRNPs
`would have to be expressed at very high levels. Perhaps the degrada-
`tion complexes act as gatekeepers, located at the nuclear pores and
`plasmodesmata, scanning the RNAs that pass through. This would
`allow mRNAs to be efficiently screened and destroyed, if recognized,
`as they exit the nucleus, thus leading to gene silencing. Viral RNAs
`
`would be scanned and destroyed as they attempt to spread from cell
`to cell through the plasmodesmata.
`
`What genes are involved in PTGS and RNAi?
`The similarity of induction, degradation and associated short
`dsRNAs in RNAi, quelling and PTGS indicates an underlying evolu-
`tionarily conserved mechanism. Analysis of mutants defective in these
`processes in Caenorhabditis elegans, Neurospora and Arabidopsis
`confirm this closeness, showing that there are a number of common
`essential enzymes or factors (Table 1). In all three species, mutation of
`an RDRP or a protein with homology to eIF2C, a rabbit protein
`thought to be involved in translation initiation42, blocks silenc-
`ing32,43–48. Another class of essential silencing proteins, those with
`homology to one of three types (RecQ, DEAH or Upf1p) of helicase,
`has been found in C. elegans, Neurospora, Chlamydomonas reinhardtii
`and Arabidopsis49–53. The roles of these proteins remain to be elucidat-
`ed. They are probably not the equivalents (or parts thereof) of Dicer-1
`in Drosophila; comparisons of the Dicer-1 sequence with genome
`databases identify K12H4.8 in C. elegans and CAF1 in Arabidopsis as
`homologues. These two Dicer-like proteins each have an RNA helicase
`domain, RNase III motifs and a PAZ domain39,54 (Fig. 3).
`There are two further categories of silencing-deficient mutants in
`plants and nematodes. One contains mutations of proteins that affect
`the structure and/or transcriptional status of chromatin, including
`DDM1, which remodels chromatin structure, MET1, which is a
`methyltransferase, and RDE2, RDE3 and MUT2, which seem to be
`involved with repressing the activity of TEs. The other category
`contains SGS3 in Arabidopsis and RDE4 from C. elegans, whose
`functions are a complete mystery. SGS3 has been cloned and
`sequenced, but has no recognizable motifs or matches with other
`sequences in available databases.
`These mutation studies show that PTGS, RNAi and quelling are
`not just the result of Dicer complexes waiting to degrade dsRNA and
`homologous ssRNA that invades a cell. Other associated processes
`are clearly involved, including a possible link to the translation
`apparatus, an RDRP, and interactions with chromosomal DNA.
`
`The role of methylation and chromatin remodelling in PTGS
`DNA methylation and chromatin structure have an integral role in
`TGS. In this form of silencing, the promoter and sometimes the
`
`836
`
`© 2001 Macmillan Magazines Ltd
`
`NATURE | VOL 411 | 14 JUNE 2001 | www.nature.com
`
`Benitec - Exhibit 1013 - page 3
`
`

`
` DEAH helicase
`
` PAZ
`
`RIII
`
`RIII
`
`dsB
`
`
`
` PAZ PAZ
`
`
`PiwiPiwi
`
`
`
`Figure 3 Distribution of domains on DCR1-like and AGO1-like proteins. Top panel
`illustrates the domains on DCR1-like proteins of ~2,000 amino acids (including
`DCR1, CAF1 and K12H4.8); bottom panel represents AGO1-like proteins of ~900
`amino acids (including AGO1, RDE1 and QDE2). RIII, RNase III domain; dsB, double-
`stranded RNA-binding domain(s). For more detailed information on domains, see
`http://www.sanger.ac.uk/Software/Pfam.
`
`coding region of the silenced transgenes are densely methylated55.
`Methylation, or methylation-associated chromatin remodelling, of
`promoter sequences is thought to prevent binding of factors neces-
`sary for transcription55. The coding sequences of PTGS-inducing
`transgenes are also frequently found to be methylated. PTGS can be
`established in plants with a mutant methyltransferase (metI), but
`during growth, the silencing becomes impaired, reactivating the
`silenced gene in sectors of the plant56. Furthermore, PTGS can
`fail to establish in mutant plants lacking the chromatin remodelling
`protein DDM1. These results suggest a role for DNA methylation
`and/or chromatin structure in both establishment and maintenance
`of PTGS.
`The mechanisms of PTGS and TGS may have more in common
`than was previously thought. In PTGS, the short RNAs derived from
`the transcribed region of the transgene act as guides for siRNPs to
`degrade target ssRNA. In TGS plants, hpRNAs containing promot-
`
`I/R transgenes
`
`Methyltransferase
`
` CH3
`
`Hairpin RNA
`
`RNA-directed
`methylation
`
`insight review articles
`
`er-region sequences are processed into short dsRNAs, and seem to
`direct methylation30. Similarly, virus-replicated RNAs direct
`sequence-specific DNA methylation57,58 and are associated with
`short dsRNAs58. It is possible that the steps of PTGS and TGS are, in
`fact, the same and differ only in their target sequences: hpRNA or
`dsRNA is cleaved by the plant homologue of Dicer-1 into
`~21-nucleotide dsRNAs to guide specific ssRNA degradation in the
`cytoplasm, and a similar ribonucleoprotein complex passes into the
`nucleus to direct chromatin remodelling/methylation of homolo-
`gous DNA (Fig. 4). Thus, production of hpRNA/dsRNA that
`contains promoter sequences leads to the methylation/altered state
`of the promoter DNA, causing TGS, whereas hpRNA/dsRNA that
`contains coding-region sequences leads to the degradation of
`homologous mRNA, causing PTGS. The methylation of coding-
`region DNA in PTGS and the potential degradation of promoter-
`sequence transcripts in TGS would be irrelevant by-products, as
`methylated coding regions are readily transcribed58,59 and promoter
`sequences are not usually transcribed.
`It seems unlikely that the DNA methylation mechanism associat-
`ed with PTGS and TGS is involved directly in protecting plants
`against most RNA viruses. The vast majority of these viruses have
`exclusively cytoplasmic
`lifecycles and no homologous DNA
`sequences in plant genomes. It is possible that dsRNA-directed
`methylation is involved in inhibiting the handful of known plant
`retroviruses or pararetroviruses during their DNA phases within the
`nucleus. It is even more likely that the mechanism is primarily for
`defence against TEs.
`
`Defence against transposons in plants
`DNA methylation may have evolved as an epigenetic means of
`containing the spread of TEs in host genomes4,60. De novo DNA methy-
`lation was first detected in plants during the inactivation of class II
`TEs61 and has been associated with both transcriptional inactivation
`
`Replicating
`RNA Virus
`
`dsRNA
`
`RNA-directed
`Dicer complex
`
`Cleavage
`
`Spread for
`systemic
`silencing
`
`Figure 4 A model for the initiation and operation of PTGS. The hpRNA or dsRNA produced from either an inverted-repeat transgene or a replicating virus is cleaved into ~21-
`nucleotide fragments by the Dicer-containing complex and used as guides for cleavage of homologous ssRNA (described in more detail in Fig. 2). The ~21-nucleotide dsRNA, in
`some sort of complex, is transported into the nucleus to direct DNA methylation of homologous sequences or into neighbouring cells to act as a mobile PTGS signal.
`
`NATURE | VOL 411 | 14 JUNE 2001 | www.nature.com
`
`© 2001 Macmillan Magazines Ltd
`
`837
`
`Benitec - Exhibit 1013 - page 4
`
`

`
`insight review articles
`
`a
`
`b
`
`c
`
`d
`
`e
`
`Figure 5 Possible ways in which transposons may generate hpRNA or dsRNA. a, LTR
`transposon integrated as an inverted repeat. b, LTR transposon integrated in the
`opposite polarity into another copy of the same transposon. c, TIR transposon
`adjacent to an endogenous promoter. d, TIR transposon (Mu). e, LTR transposon
`adjacent to an endogenous promoter. Red blocks represent either the inverted or
`direct terminal-repeat sequences. Green blocks represent the coding regions of the
`TEs. Dark blue arrows below represent the transcripts generating dsRNAs or hpRNAs,
`and the drawing below each transcript represents the structure it may form (that is,
`either a hairpin or a dsRNA structure). Blue box represents the transcription terminator
`sequence and light blue arrows represent readthrough transcription. Large green
`arrows represent endogenous promoters.
`
`and increased transitional mutation62–64. Many studies have shown the
`involvement of methylation with transposon inactivation61,65–67 and
`demethylation with transposon activation68,69. A recent demonstration
`of this comes from the study of the retrotransposon Tto1 in Arabidop-
`sis70. As this TE becomes transcriptionally silenced, it also becomes
`increasingly methylated, and demethylation of the element, in a
`hypomethylation mutant ddm1 background, reactivates its transcrip-
`tion. However, it is still not entirely clear whether the methylation
`itself inactivates the transposon or whether it is a secondary effect of
`inactivation caused by a change in chromatin structure.
`Epigenetic inactivation of TEs occurs in many, possibly most,
`cases by its insertion into or near an already heterochromatic block of
`genomic DNA and the radiation of this repressed state into the
`elements3. But TEs integrating into euchromatic areas may well be
`the target for dsRNA-induced silencing. There are a number of
`scenarios (Fig. 5) of how TEs could produce dsRNA or hpRNA to
`trigger this mechanism. The LTRs of class I TEs contain promoter
`sequences, so two TE copies integrating as an inverted repeat could
`produce hpRNA transcripts of these sequences. TEs often integrate
`within each other, potentially generating transcribable, complex
`inverted-repeat sequences. Some transposons, such as Robertson’s
`mutator (Mu), have convergently arranged genes which produce
`transcripts that, by failing to terminate or be polyadenylated at the
`appropriate sequences, have regions of complementarity with each
`other71. Insertion of a class II TE adjacent to an endogenous promoter
`directing transcription across the elements could produce RNA with
`self-complementarity from the TIRs. An adjacent endogenous
`promoter directing transcription from the reverse end of a TE could
`produce antisense RNA that might hybridize with the TE RNA to
`produce dsRNA. It is also possible that the reverse transcription of
`
`retrotransposon RNAs produces intermediates in the cytoplasm
`similar to replicating RNA viruses which, although RNA/DNA
`hybrids rather than RNA/RNA hybrids, act as triggers for dsRNA-
`induced silencing. Indeed, normal infection of plants with cauli-
`flower mosaic pararetrovirus, which will produce a similar
`RNA/DNA intermediate, triggers a PTGS-like response72. But the
`most convincing evidence that the dsRNA-induced silencing
`mechanism is suppressing TEs is that such silenced elements are
`reactivated in a number of PTGS- and RNAi-defective mutants
`(Table 1), and that some of the ~21-nucleotide dsRNAs from RNAi
`and PTGS extracts contain sequences of TEs (ref. 40, and A. J. Hamil-
`ton and D. C. Baulcombe, personal communication). The TEs are
`probably controlled by methylation of their DNA and degradation of
`their transposase mRNA.
`
`PTGS can spread systemically through a plant
`PTGS has three phases: initiation, maintenance and, remarkably,
`spread16,73. Transgenes and viruses can initiate PTGS, as can
`exogenous DNA delivered by bombardment or Agrobacterium
`infiltration16, and grafting of unsilenced scions onto silenced
`rootstock73. These last methods give localized delivery points for
`PTGS that spreads from these points into other tissues. It seems to
`spread by a non-metabolic, gene-specific diffusible signal that is
`capable of travelling both between cells through plasmodesmata, and
`long distances via the phloem16,73. For example, new tissue
`growing from a GUS-expressing scion, grafted onto a GUS-silenced
`rootstock, shows progressive silencing of its GUS transgene73. The
`signal seems to be sequence specific, to move uni-directionally
`from source to sink tissue, and can traverse at least 30 cm of wild-type
`stem grafted between the GUS-expressing scion and GUS-silenced
`rootstock73.
`To account for the specificity of the signal, it must consist (at least
`in part) of the transgene product, probably in the form of RNA73. The
`concept of cell-to-cell and long-distance spread of endogenous RNAs
`within plants remains somewhat controversial, but is not unprece-
`dented. For instance, plant viruses have genomes composed of RNA
`and, when they infect their host, their RNA spreads throughout the
`plant. Viral-encoded movement proteins facilitate the movement of
`viral RNA between cells through plasmodesmata in the form of
`either a ribonucleoprotein complex or intact virions. To fulfil
`this role, movement proteins have the capacity to move between
`cells, bind viral RNA and dilate the size exclusion limit of plasmodes-
`mata74. Simpler still, viroids — plant pathogens with small
`(~350 nucleotide) naked RNA genomes encoding no proteins — also
`infect and spread though plants, presumably associated with
`host proteins75.
`There is an emerging picture of RNA mobility in plants that
`potentially impacts on other plant processes, including transport of
`the gene-specific silencing signal. Examples of host RNAs moving
`from cell to cell include the KNOTTED1 transcription factor and its
`corresponding mRNA76, and the transcript that encodes sucrose
`transporter 1, which has been localized to the enucleate sieve
`elements, presumably having been transported there from the asso-
`ciated companion cell77. Perhaps the most convincing demonstra-
`tion of intercellular movement of endogenous plant RNA, and
`potentially signalling, is the demonstration that mRNA is found in
`the phloem of rice and cucurbits78,79. Mobility of pumpkin phloem
`RNA was demonstrated using grafting experiments. In one instance,
`a transcript encoding a transcription factor, NACP, was detected in
`the meristem of cucumber scions that had been heterografted onto a
`pumpkin rootstock78. Thus RNA molecules, derived from the
`silenced transgene, might move from cells where this gene is silenced,
`possibly with cellular protein factors fulfilling a role similar to viral
`movement proteins, to induce silencing in other cells expressing the
`same transgene. This raises three questions — is the signal the
`~21-nucleotide dsRNA, how is the signal propagated, and what is the
`natural (non-transgenic) role of the signal?
`
`838
`
`© 2001 Macmillan Magazines Ltd
`
`NATURE | VOL 411 | 14 JUNE 2001 | www.nature.com
`
`Benitec - Exhibit 1013 - page 5
`
`

`
`Figure 6 Tobacco plants showing potato virus Y (PVY)
`susceptibility, immunity and resistant/recovery
`symptoms. a, Non-transgenic tobacco plant challenged
`with PVY, showing a uniform chlorotic mosaic on leaves
`throughout the plant. b, Virus-challenged transgenic
`plant containing a transgene that expresses a hpRNA
`derived from PVY sequences, showing immunity to the
`virus and no symptoms. c, A whole leaf, and d, close-
`up, from a PVY-challenged transgenic plant showing
`resistance/recovery symptoms. As the plant grows the
`leaves show fewer and fewer yellow patches; only the
`yellow patches contain detectable levels of virus. The
`pattern suggests that as the virus attempts to spread
`from the phloem into the surrounding cells, a signal is
`emitted that allows more distal cells to be forewarned
`and resist and restrict the virus.
`
`a
`
`c
`
`insight review articles
`
`b
`
`d
`
`Defence is a two-step process
`Most plant species are immune to most plant viruses, possibly due to
`compatibility requirements between the RDRP or movement
`protein of a specific virus and the host factors present in the plant.
`However, it is also possible that the only virus/host combinations
`leading to an infection are ones in which the virus can overcome or
`avoid the plant’s PTGS defence mechanism. For example, potyvirus-
`es encode a protein, HC-Pro, that potently disables PTGS31,80,81 by
`directly or indirectly preventing the dsRNA cleaving activity of the
`Dicer complex. Therefore, a plant challenged with a potyvirus
`becomes a battleground in the fight between a defence and a counter-
`defence strategy. It is a race between the cell’s recognition and
`degradation of viral RNA, and the virus expressing HC-Pro to
`inactivate the degradation machinery. Perhaps the mobile signal,
`seen in the artificial situation of transgenes and grafting experiments,
`is a reflection of a plant’s fallback strategy. If the first-challenged cells
`ar

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket