Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3889 2023-07-06 13:01:34 |
2 format Meta information modification 3889 2023-07-07 04:17:14 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Vermeulen, A.; Takken, F.L.W.; Sánchez-Camargo, V.A. Translation Arrest. Encyclopedia. Available online: https://encyclopedia.pub/entry/46520 (accessed on 03 September 2024).
Vermeulen A, Takken FLW, Sánchez-Camargo VA. Translation Arrest. Encyclopedia. Available at: https://encyclopedia.pub/entry/46520. Accessed September 03, 2024.
Vermeulen, Annemarie, Frank L. W. Takken, Victor A. Sánchez-Camargo. "Translation Arrest" Encyclopedia, https://encyclopedia.pub/entry/46520 (accessed September 03, 2024).
Vermeulen, A., Takken, F.L.W., & Sánchez-Camargo, V.A. (2023, July 06). Translation Arrest. In Encyclopedia. https://encyclopedia.pub/entry/46520
Vermeulen, Annemarie, et al. "Translation Arrest." Encyclopedia. Web. 06 July, 2023.
Translation Arrest
Edit

Plants evolved several mechanisms to protect themselves against viruses. Besides recessive resistance, where compatible host factors required for viral proliferation are absent or incompatible, there are (at least) two types of inducible antiviral immunity: RNA silencing (RNAi) and immune responses mounted upon activation of nucleotide-binding domain leucine-rich repeat (NLR) receptors. RNAi is associated with viral symptom recovery through translational repression and transcript degradation following recognition of viral double-stranded RNA produced during infection. NLR-mediated immunity is induced upon (in)direct recognition of a viral protein by an NLR receptor, triggering either a hypersensitive response (HR) or an extreme resistance response (ER).

antiviral resistance NLR-mediated translational arrest PTGS

1. Translation of Viral Transcripts in Plants

1.1. Adaptations of Viral RNAs for Enhanced Translation

Viral RNAs must disguise themselves to mimic eukaryotic mRNA, evade the host immune system, and hijack the host machinery for translation. Eukaryotic mRNA has a unique 5′ cap structure essential for translation initiation. The cap structure, which consists of a methylated guanosine residue, binds translation factors and ribosomes to initiate protein synthesis. Additionally, eukaryotic mRNA has a 3′ poly(A) tail that protects the mRNA from degradation by exonucleases [1]. Many plant RNA viruses lack either a 5′ cap, a 3′ poly(A) tail, or both, preventing recognition and translation by the host machinery. Moreover, many plant RNA viruses are multicistronic, containing several open reading frames (ORFs) within the same RNA strand or encoding a polypeptide that is processed by proteases into multiple peptides. To overcome the lack of a 5′ cap or a 3′ poly(A) tail, viral RNAs have evolved adaptations to recruit ribosomes. These adaptations include, e.g., 5′ internal ribosome entry sites (IRESs), 5′ viral genome-linked protein (VPg), 3′ cap-independent translation enhancers (CITEs), and cap snatching [2]. IRES elements are highly structured RNA sequences that allow ribosomes to initiate translation from a position internal to the RNA molecule, bypassing the requirement of a 5′ cap structure. VPg is a protein covalently linked to the 5′ end of some viral RNAs, and it is thought to enhance translation initiation [2]. CITEs are RNA elements located at the 3′ end of some viral RNAs that enhance translation initiation by interacting with translation initiation factors. Interestingly, both 5′ IRES and 3′ CITE elements are also present in some eukaryotic mRNAs and are thought to stimulate translation during stress and developmental circumstances when cap recognition is hindered [3]. These elements consist of cis-acting secondary structures that recruit host initiation factors or ribosomal subunits, enabling protein synthesis [4][5][6]. Another strategy involves VPg, which plays a crucial role in the translation of viral RNA by interacting with host translation factors, such as eIF4E and its isoform eIF4isoE. VPg competes with the cap structure of host mRNAs to bind these factors, thereby inhibiting host translation and redirecting the host machinery towards viral translation [7][8]. In contrast, some segmented negative-stranded RNA viruses have developed an alternative mechanism called cap snatching to initiate translation [9][10].

1.2. Viral Translation

Different RNA virus families have developed diverse mechanisms to achieve efficient translation. Depending on the location of the encoded peptide sequence in the viral RNA, commonly used strategies include cap-independent translation, synthesis of subgenomic RNAs (sgRNAs), cap snatching, and translational recoding [11][12][13]. Cap-independent translation occurs at the first 5′ localized ORF and is regulated by IRES and 3′ CITE structures in the viral RNA. These structures interact with translation initiation factors, such as eIF4E and eIF4G, or directly with the ribosomal subunit through 18s rRNA, which stimulates host translation initiation complex assembly and translation of the viral RNA, mimicking cap-dependent translation of eukaryotic mRNA [4][5][6].
In positive-stranded RNA viruses, sgRNAs are synthesized from initial viral RNA by an RNA-dependent RNA polymerase (RdRp) encoded in the viral genome. RdRp recognizes subgenomic promoters in the viral RNA that give rise to different sgRNAs. These sgRNAs enable the host machinery to translate viral proteins located internally or at the 3′ end of the viral RNA [11]. Due to the dense coding of viral genomes, overlapping or adjacent ORFs often require translational recoding for translation. Translational recoding includes leaky ribosome scanning, non-AUG initiation, ribosomal codon read-through, ribosomal frameshifts, and translational bypassing. The mechanism used largely depends on the viral genus [14].

1.3. Optimization of Viral Translation

In addition to cis elements that regulate translation, viruses have evolved various strategies to achieve maximum translation efficiency. For example, a commonly observed phenomenon is the induction of a ‘host shut-off’ mechanism, where the translation of endogenous mRNA is suppressed [15]. This can be accomplished through interference with the cap-dependent translation of host mRNAs, resulting in a decrease in, among other things, host antiviral responses and an increase in viral RNA translation. Moreover, viruses have optimized translation by facilitating RNA cyclization, interfering with host translation initiation, and compartmentalizing translation in virus factories (VFs) [16][17]. Viral RNA cyclization is promoted through a specific type of 3′ CITE structure that binds to eIF4F, leading to the interaction of eIF4F with a hairpin structure at the 5′ end of the RNA. In turn, this results in the cyclization of translation [18][19]. This cis-element-stimulated cyclization has been demonstrated for viruses in the Tombusviridae family [18][19]. Multiple viruses interfere with host translation by targeting eIF4E, altering its phosphorylation status [20]. In plants, phosphorylated eIF4isoE shows an increased binding affinity for VPg and enhanced mRNA translation [21][22], perhaps favoring the translation of viral VPg-containing RNAs over host mRNA. VFs, viroplasms, inclusion bodies (IBs), or viral replication complexes (VRCs) are intracellular structures induced by viruses. These membrane-bound inclusion-like bodies or spherules concentrate viral RNA and proteins, potentially enhancing viral replication. The formation of VFs is often facilitated by viral proteins that manipulate membranes of the endoplasmic reticulum, mitochondria, peroxisomes, and/or chloroplast membranes [23][24][25][26][27]. Although VFs are primarily associated with RNA replication, they have also been suggested to play a role in viral translation and cell-to-cell movement in certain cases [28][29].

2. Immune Mechanisms Resulting in Translational Repression of Viral RNAs

2.1. Viral Recovery through PTGS-Mediated Translational Repression

Symptom recovery is observed in certain plant–virus interactions, where asymptotic leaves emerge or symptomatic leaves recover after systemic infection with a virus. This phenomenon has been reported for numerous unrelated viruses and can also be induced through mutation of VSRs [30][31][32][33][34]. Symptom recovery is associated with systemic sequence-specific resistance, providing protection against reinfections and cross-protection against related viruses [33]. Symptom recovery from RNA viruses can result from either degradation or translational repression of viral transcripts [35][36][37][38][39]. Symptom recovery is the only known phenomenon in plants where PTGS leads to translational repression [30][31].
In Nicotiana benthamiana, symptom recovery from tomato ringspot virus (ToRSV) is associated with PTGS but not accompanied by viral clearance [37]. Instead, ToRSV recovery is accompanied by an accumulation of the ToRSV RNA2 viral segment and a reduction in its encoded CP and MP through translational repression [35][38]. Whether translation of RNA1 of ToRSV is also repressed remains elusive. Similarly, symptomless recovery from an engineered tobacco rattle virus clone (TRV-GFP) infection in Arabidopsis thaliana is associated with a decrease in GFP fluorescence, followed by a drop in the viral titer [36][39]. During this stage, the TRV RNA is less associated with ribosomes, which results in reduced levels of GFP. This observation suggests that translational repression is involved in TRV recovery [36]. In contrast, in N. benthamiana, TRV RNA is known to be targeted for slicing [40]. In A. thaliana undergoing infection by oil-seed rape mosaic virus (ORMV), symptom recovery occurs in newly developed leaves in the presence of infection-competent replicating viral RNAs. It has been proposed that 21–22 nt siRNAs in concert with PTGS and TGS machinery are responsible for this recovery in a non-cell-autonomous manner, generating a source-to-sink siRNA gradient, where VSR function becomes saturated, perhaps inhibiting viral protein translation. Accordingly, A. thaliana mutants that result in a dysfunctional PTGS and/or TGS fail to undergo symptom recovery, while impairing RNA decay machinery enhances such a response, probably by allowing the accumulation of dsRNA, which can be processed into vsRNAs and exported to new leaves [41]. These data suggest that PTGS participates in translational repression during viral recovery. Moreover, under certain interactions, translational repression and slicing can occur in parallel [38].

2.2. NLR-Mediated Translational Arrest

During NLR-mediated TA, the translation of viral transcripts of the virus triggering the immune response is arrested, as well as that of transcripts from other viruses present in the cell [42][43][44]. The resulting TA appears to be cell-autonomous and only occurs in cells where the NLR is activated [42][43]. Two structurally divergent NLRs, a CNL (Rx1) and a TNL (N), are known to induce TA in N. benthamiana [42][45].
Rx1 has been introgressed into commercial potato cultivars from the wild potato species Solanum andigena [46][47][48]. The Rx1 gene confers extreme resistance to PVX. In transgenic N. benthamiana, Rx1 is activated in the cytoplasm upon recognition of the PVX CP [47][49][50][51][52]. Recently, it was shown that Ran GTPase Activating Proteins 1/2 (RanGAP1/2) are targets of PVX CP [53]. Since inactive Rx1 and RanGAP2 interact through their CC and WPP domains, respectively, Rx1 activation likely occurs by indirect recognition through RanGAP2 [53][54][55]. Once activated, Rx1 is translocated to the nucleus, which is essential to the mounting of a full immune response [45][51][52][56]. Rx1 shuttling is mediated by RanGAP2 and the co-chaperone SUPRESSOR OF G2 ALLELE OF SKP1 (SGT1) [52][55][56]. In the nucleus, the activated form of Rx1 directly binds and distorts double-stranded DNA through its NB-ARC domain [49]. The DNA-binding capacity of Rx1 could enable it to function as a transcriptional regulator by facilitating access to DNA for transcription machinery. The binding specificity of Rx1 is suggested to be provided by the Golden2-like transcription factor (TF) NbGlk1 [57]. In the absence of PVX, the CC domain of Rx1 binds to NbGlk1 and a bromodomain (BD)-containing protein (NbBDCP), preventing chromatin interaction [58]. Upon PVX infection, NbGlk1 mediates Rx1 binding to Golden2-like consensus DNA sequences, which may lead to TA by regulating unknown target genes.
N mediates resistance to TMV through recognition of the p50 helicase domain present in the TMV replicase [59][60]. In N. tabacum, activation of this NLR results in HR [61]. However, when introduced in N. benthamiana, N activation leads to ER through TA [42][61]. Recognition of p50 is mediated by N-receptor-interacting protein 1 (NRIP1), which is normally localized in chloroplasts [62]. However, p50 recruits NRIP1 to the cytoplasm and nucleus to form p50-NRIP1 complexes [62]. Afterward, NRIP1 binds to the TIR domain of N, triggering activation [62][63]. The N-mediated defense response encompasses multiple pathways. However, the specific mechanism by which it induces TA remains unknown. Currently, the only protein suggested to be involved in TA is the helper NLR N REQUIREMENT GENE 1 (NRG1) via an unknown mechanism [43]. Future studies aimed at unveiling the proteins involved in NLR-mediated transcriptional activation could help determine whether both NLRs require the same partners for signaling or utilize distinct pathways. Differences in sensitivity to the VSR p38 of turnip crinkle virus (TCV) suggest that the mechanisms that lead to TA are different for N and Rx1 [45].
Alternatively, virus-derived nucleic acids from begomovirus, a DNA virus, can trigger a global TA through NIK1 [64][65]. NIK1 is part of the same LRR-RLK subfamily as BRI1-ASSOCIATED RECEPTOR KINASE-1 (BAK1), which is well known for its role in plant defense against bacteria, fungi, and oomycetes [66][67][68]. The global TA following NIK1 activation is thought to be induced through the downregulation of translational-machinery-related genes [69]. This downregulation is indirectly facilitated by the ribosomal protein L10A (RPL10A), which translocates to the nucleus upon phosphorylation by NIK1, where it interacts with the L10-interacting MYB domain-containing (LIMYB) TF [65][65]. However, the mechanism of NIK1 activation and the signaling pathway leading to global TA remain obscure, though transcriptional reprogramming might be common during TA in response to plant viruses.

3. Translational Repression

3.1. Transcript Recognition

To mount a TA of viral RNAs, the plant must distinguish viral RNAs from endogenous RNAs. Recognition of a viral transcript can occur through the sequence specificity brought by vsRNA in an RISC complex or by the binding of RNA-binding proteins (RBPs) to common viral motifs present in the RNA. Some of these common motifs include dsRNA, the 5′-ppp region of the RNA, the 5′ UTR region, and hairpin structures [70][71][72][73].
During viral recovery, transcript recognition is assumed to occur through sequence specificity provided by siRNAs [74]. This hypothesis is supported by the observation that plants have sequence-specific resistance to reinfection after viral recovery [33]. However, transcript recognition has not been reported for ToRSV or TRV recovery, where translational repression is involved. In NLR-mediated TA, viral transcripts present in the cell are halted [42][43][44]. Most likely, a common element present in viral RNA is recognized by an RBP, resulting in the translational repression of viral transcripts. Evidence suggests that a structural motif present in the CP-encoding part of the PVX RNA is essential for TA of these transcripts [42]. In correspondence with this, the translation of PVX transcripts lacking the CP region was no longer halted when N-mediated TA was triggered. The PVX-induced TA could be reverted when the CP ORF was replaced with another viral CP with low sequence similarity, suggesting that the recognition and subsequent targeting for repression is likely caused by a structural element in this region. Moreover, activation of the NLR Tm-2a from a wild relative of tomato (Solanum lycopersicum), Solanum peruvianum, inhibited the accumulation of PVX but not that of PVXΔCP in tobacco [42]. This suggests that the structural element present in the CP region might be a common target for NLRs. However, the possible binding partner and RNA structure both remain unidentified. Other elements of the viral RNAs, such as their 5′ cap structure and their poly-A tail at the 3′ region, are unlikely to be recognized during NLR-mediated TA, since TCV sgRNA lacks these structures and is still halted during N-mediated TA [43][75].

3.2. Translation Inhibition

After dsRNA sensors recognize the viral transcripts, the translation of transcripts can either be repressed before translation initiation or later, during translation elongation. For example, in miRNA-mediated translational repression, translation initiation is known to be inhibited through the physical blocking of mRNAs by AGO1-RISC binding at the 5′ UTR to prevent recruitment of the 48S ribosomal subunit and ribosome assembly [76]. In addition, depending on the position of the RISC target site, the binding of RISC to the RNA can physically block translation elongation [76]. The stage during translation at which the transcripts are halted can be determined through polysome profiling, as the association of ribosomes is prevented by translational inhibition prior to translation initiation [77]. During ToRSV recovery, monosome and polysome profiling revealed that translation inhibition occurred after translation initiation. However, during TRV recovery, translation inhibition appeared to occur before translation initiation [36][38]. In NLR-mediated TA, polysome profiles revealed that PVX RNAs were no longer associated with polysomes [43]. In N-mediated TA, PVX RNA was prohibited from associating with monosomes, which suggests that repression occurs before translation initiation [42]. However, the mechanism behind translation initiation inhibition during NLR-mediated TA remains obscure.

3.3. Processing Bodies

Liquid–liquid phase separation (LLPS) is a physical phenomenon where a homogeneous solution separates into two distinct liquid phases, driven by changes in concentration, temperature, or other factors [78][79]. RNA granules, which comprise all intracellular aggregations of ribonucleoprotein (RNP) complexes large enough to be microscopically visible, can undergo LLPS. Cells undergoing a viral infection often accumulate RNA granules, which dynamically influence each other by exchanging components, such as transcripts and proteins [80].
In plants, RNA granules can be classified into PBs, stress granules (SGs), and siRNA bodies. In PBs and SGs, LLPS behavior is thought to be driven by multivalent interactions between RBPs, RNA, and other proteins that enable the formation of weak reversible bonds, which, in turn, can dynamically change the properties of the granules [81][82][83][84]. PBs are essential in regulating gene expression, mRNA decay, and translation [85]. These bodies reside within cells to process aberrant mRNA transcripts. During mRNA decay, endogenous aberrant mRNA or viral RNAs are degraded through decapping by the mRNA decapping protein 2 (DCP2), followed by de-adenylation and 5′ to 3′ decay by the exoribonuclease XRN4 [85]. Decapping is assisted by the co-activator DCP1 and scaffold varicose (VCS) [86][87]. Besides mRNA decay, PBs are sites for storing translationally repressed RNAs [82]. Once the abundance of repressed RNA increases due to RNAi, virus infection, or UV irradiation, the number and size of PBs changes consequently [36][43][85][88][89]. Plant viruses have been proposed to utilize components of PBs for their benefit, and PBs have been suggested to store repressed viral RNA during immune responses and viral recovery [36][43][90][91].
During TRV recovery, A. thaliana shows an increase in PBs potentially containing repressed TRV RNA aggregates. Eventually, these RNAs are degraded by decapping enzymes present in these PBs. However, degradation of the TRV RNA is not necessary for recovery, as was demonstrated by the normal TRV recovery observed in the DCP2 mutant its1 [36]. Although DCP1 was used as a PB marker, the exact composition of the PBs that form during TRV recovery remains unknown.
When cells underwent NLR-mediated TA, PBs were shown to accumulate as well [43]. These PBs contained DCP1 but were depleted of DCP2, resulting in transcript accumulation caused by the lack of RNA degradation [43]. The protein composition within these types of PBs remains unknown so far. After NLR activation, the cellular DCP1 levels remain unchanged, which suggests that a pre-existing pool of DCP1 acts as a nucleation point for PB formation [43]. While Meteignier et al. (2016) suggested that the lack of DCP2 in PBs could be due to limited levels of cellular DCP2, previous studies have shown that DCP1 and -2 are both recruited into PBs depending on the stress perceived and that this is not necessarily reflective of their cellular concentrations [92]. Therefore, it is likely that the depletion of DCP2 in NLR-mediated PB formation is an active exclusion rather than a passive reflection of PB protein component concentrations in the cytoplasm. However, the mechanisms that stimulate PB formation during NLR-mediated immunity remain to be elucidated.
The formation of PBs in response to NLR activation occurs through a different mechanism than that triggered by UV irradiation or RNAi [43][88]. For example, UV irradiation triggers phosphorylation of eIF2a, resulting in a global TA [93], which does not occur after N activation [43]. Additionally, when PTGS is repressed by VSR P19, PB formation is reduced, while P19 has no influence on PB formation following N activation [43]. Although the mechanisms that induce PB formation after NLR activation are distinct, whether the formed PBs are qualitatively similar or distinct remains under debate and requires further research. Whether PBs form during ToRSV recovery is unknown. However, it is unlikely, since the inhibition of translation initiation is known to result in PB formation, whilst inhibiting translation elongation leads to a decrease in PBs [88]. As discussed, translational repression during ToRSV recovery likely happens during elongation, whilst TRV recovery and NLR-mediated TA occurs before translation initiation. However, the exact mechanisms underlying PB formation have not been investigated for these viruses.

3.4. Translation Repression Depends on AGOs and VSR Interference

AGOs form an integral part of the RNAi pathway, where they act as the catalytic subunit of the RISC complex. Multiple AGOs (AGO1, AGO2, AGO4, and AGO7) are known to be involved in and are essential for antiviral immune responses [36][42][94][95][96][97][98][99]. All these AGO proteins have RNA slicing activity, whilst only AGO1 has been proven to be directly involved in translational repression [74][75][76][94][100][101]. AGOs can directly perform RNA slicing through their endonucleolytic PIWI domain [102]. Although the composition of the AGO protein interactome remains elusive, translational repression by AGO is thought to operate through association with WG/GW motif-containing proteins [103][104][105][106]. In A. thaliana, genetic studies identified a gene encoding the GW-containing protein SUO involved in miRNA-mediated translational repression [101]. However, whether SUO directly interacts with AGO through its GW motifs to perform its function has not been investigated, and how AGOs determine whether to perform slicing or repression is poorly understood. Possibly, AGO function is regulated through post-translational modifications, such as phosphorylation or changes in subcellular localization [35].
Multiple VSRs interact with AGOs to impair RNAi. For instance, AGO1 expression, stability, and activity can be targeted by VSRs [38][100][107][108][109][110]. In line with this, VSRs, including the carmovirus p38 and ipomovirus P1, contain the AGO-interacting motif WG/GW and compete with host proteins for AGO binding [109][111]. Some VSRs have multiple strategies to interfere with plant signaling. For example, p38 of TCV prevents the processing of dsRNA into siRNA and impairs siRNA loading into AGO1 and AGO2 in A. thaliana [109][112][113].
During symptom recovery of ToRSV in N. benthamiana, translational repression is AGO1-dependent [35]. It is likely that AGO1 is (in)directly involved in the PTGS mechanism leading to translational repression of ToRSV RNA2. This is supported by the finding that ToRSV CP hinders AGO1 function through interaction with the WG/GW motif of AGO1, triggering AGO1 degradation and probably competing with cellular WG/GW proteins involved in translational repression [38][114]. During TRV infection in A. thaliana, AGO2 and AGO4 are involved in initial susceptibility to TRV, whilst other unidentified proteins are required during recovery [36]. Multiple VSRs are known to affect symptom recovery and the expression of strong VSRs, such as HC-Pro from potyvirus, and p25 from PVX can eliminate symptom recovery during ToRSV infection [115]. Similarly, viral recovery of TRV was abolished by p38 of TCV [36]. Moreover, the inactivation of VSRs can lead to symptom recovery of virus isolates that typically do not display recovery. This has been shown for 2b of cucumovirus, HC-Pro of potyvirus, and P19 of tombusvirus [116][117][118][119]. Recovery in ToRSV and TRV is not dependent on a lack of VSRs in the virus isolates but rather on the relative strength of the VSRs present. As mentioned earlier, ToRSV CP acts as an AGO-hook triggering AGO1 degradation, and TRV 16K acts as a VSR by preventing AGO4-RISC assembly [38][120]. Additionally, the loss-of-function mutants for AGO1 and other genes involved in siRNA signal amplification, such as RDR6, SGS3, and DCL4, fail to alleviate infection symptoms [41]. These observations together suggest that AGO protein(s), in complex with partners containing WG/GW motifs, could participate in the TA of RNA viruses and that VSRs target this function.
Rx1-mediated and N-mediated TA are AGO1-independent and AGO4-dependent [42]. AGO4 is involved in resistance to many viruses through the RNA-directed DNA methylation (RdDM) pathway, yet it is also involved in antiviral mechanisms unrelated to RdDM in the cytoplasm [98][121][122]. During plantago asiatica mosaic virus (PIAMV) infection, AGO4 is suggested to re-localize to the cytoplasm and directly target the PIAMV RNA [123]. The exact mechanisms behind this interaction remain unknown. It is likely that AGO4 acts independently of the RdDM pathway as well during NLR-mediated TA, since the viral replication cycle is restricted to the cytoplasm. Involvement of the AGO proteins in NLR-mediated TA is further proven by the expression of VSR p0 from beet western yellows virus (BWYV), p38 of TCV, and p19 of cymbidium ringspot virus (CymRSV) [42][45][124]. p0 inhibits N-mediated TA and is known to induce the degradation of AGO proteins in A. thaliana [42][111]. However, its effect on Rx1-mediated TA has not been tested. Possibly, N-mediated TA is repressed by p0 promoting the degradation of AGO4. Besides p0, p38 of TCV inhibits N-mediated TA [42]. However, p38 does not physically interact with AtAGO4, suggesting that N-mediated TA requires additional AGO proteins or that p38 inhibits TA at the dsRNA processing level [109]. The latter is unlikely since NLR-mediated transcript recognition is supposed to occur through common motif recognition rather than through siRNA complementarity, suggesting that the process might be siRNA-independent. Rx1-mediated TA is not inhibited by p38, suggesting a distinct mechanism leading to Rx1-mediated TA [45]. p19 of CymRSV is known to repress systemic PTGS by binding dsRNAs in order to prevent incorporation into RISC complexes. However, TA induced by both Rx1 and N is not influenced by p19, further supporting a distinction between the TA mechanisms involved in NLR and viral recovery [45][96][124].

References

  1. Passmore, L.A.; Coller, J. Roles of MRNA Poly(A) Tails in Regulation of Eukaryotic Gene Expression. Nat. Rev. Mol. Cell Biol. 2022, 23, 93–106.
  2. Geng, G.; Wang, D.; Liu, Z.; Wang, Y.; Zhu, M.; Cao, X.; Yu, C.; Yuan, X. Translation of Plant RNA Viruses. Viruses 2021, 13, 2499.
  3. Weingarten-Gabbay, S.; Elias-Kirma, S.; Nir, R.; Gritsenko, A.A.; Stern-Ginossar, N.; Yakhini, Z.; Weinberger, A.; Segal, E. Comparative Genetics: Systematic Discovery of Cap-Independent Translation Sequences in Human and Viral Genomes. Science 2016, 351, aad4939.
  4. Jang, S.K.; Krausslich, H.-G.; Nicklin, M.J.H.; Duke, G.M.; Palmenberg, A.C.; Wimmer’, E. A Segment of the 5′ Nontranslated Region of Encephalomyocarditis Virus RNA Directs Internal Entry of Ribosomes during in Vitro Translation. J. Virol. 1988, 62, 2636–2643.
  5. Karetnikov, A.; Lehto, K. The RNA2 5′ Leader of Blackcurrant Reversion Virus Mediates Effecient in Vivo Translation through an Internal Ribosomal Entry Site Mechanism. J. Gen. Virol. 2007, 88, 286–297.
  6. Roberts, R.; Mayberry, L.K.; Browning, K.S.; Rakotondrafara, A.M. The Triticum Mosaic Virus 5′ Leader Binds to Both EIF4G and EIFiso4G for Translation. PLoS ONE 2017, 12, e0169602.
  7. Khan, M.A.; Miyoshi, H.; Gallie, D.R.; Goss, D.J. Potyvirus Genome-Linked Protein, VPg, Directly Affects Wheat Germ in Vitro Translation: Interactions with Translation Initiation Factors EIF4F and EIFiso4F. J. Biol. Chem. 2008, 283, 1340–1349.
  8. Miyoshi, H.; Okade, H.; Muto, S.; Suehiro, N.; Nakashima, H.; Tomoo, K.; Natsuaki, T. Turnip Mosaic Virus VPg Interacts with Arabidopsis thaliana EIF(Iso)4E and Inhibits in Vitro Translation. Biochimie 2008, 90, 1427–1434.
  9. Duijsings, D.; Kormelink, R.; Goldbach, R. Alfalfa Mosaic Virus RNAs Serve as Cap Donors for Tomato Spotted Wilt Virus Transcription during Coinfection of Nicotiana benthamiana. J. Virol. 1999, 73, 5172–5175.
  10. Estabrook, E.M.; Tsai, J.; Falk, B.W. In Vivo Transfer of Barley Stripe Mosaic Hordeivirus Ribonucleotides to the 5′ Terminus of Maize Stripe Tenuivirus RNAs. Proc. Natl. Acad. Sci. USA 1998, 95, 8304–8309.
  11. Kneller, E.L.P.; Rakotondrafara, A.M.; Miller, W.A. Cap-Independent Translation of Plant Viral RNAs. Virus Res. 2006, 119, 63–75.
  12. Miller, W.A.; Koev, G. Synthesis of Subgenomic RNAs by Positive-Strand RNA Viruses. Virology 2000, 273, 1–8.
  13. Rodnina, M.V.; Korniy, N.; Klimova, M.; Karki, P.; Peng, B.Z.; Senyushkina, T.; Belardinelli, R.; Maracci, C.; Wohlgemuth, I.; Samatova, E.; et al. Translational Recoding: Canonical Translation Mechanisms Reinterpreted. Nucleic Acids Res. 2020, 48, 1056–1067.
  14. Dever, T.E.; Dinman, J.D.; Green, R. Translation Elongation and Recoding in Eukaryotes. Cold Spring Harb. Perspect. Biol. 2018, 10, a032649.
  15. Wang, D.; Maule, A.J. Inhibition of Host Gene Expression Associated with Plant Virus Replication. Science 1995, 267, 229–231.
  16. Jan, E.; Mohr, I.; Walsh, D. A Cap-to-Tail Guide to MRNA Translation Strategies in Virus-Infected Cells. Annu. Rev. Virol. 2016, 3, 283–307.
  17. Fernández de Castro, I.; Tenorio, R.; Risco, C. Virus Factories. In Encyclopedia of Virology; Bamford, D.A., Zuckerman, M., Eds.; Academic Press: San Diego, CA, USA, 2021; pp. 495–500.
  18. Miras, M.; Truniger, V.; Querol-Audi, J.; Aranda, M.A. Analysis of the Interacting Partners EIF4F and 3′-CITE Required for Melon Necrotic Spot Virus Cap-Independent Translation. Mol. Plant Pathol. 2017, 18, 635–648.
  19. Nicholson, B.L.; Wu, B.; Chevtchenko, I.; White, K.A. Tombusvirus Recruitment of Host Translational Machinery via the 3′ UTR. RNA 2010, 16, 1402–1419.
  20. Royall, E.; Doyle, N.; Abdul-Wahab, A.; Emmott, E.; Morley, S.J.; Goodfellow, I.; Roberts, L.O.; Locker, N. Murine Norovirus 1 (MNV1) Replication Induces Translational Control of the Host by Regulating EIF4E Activity during Infection. J. Biol. Chem. 2015, 290, 4748–4758.
  21. Khan, M.A. Phosphorylation of Translation Initiation Factor EIFiso4E Promotes Translation through Enhanced Binding to Potyvirus VPg. J. Biochem. 2019, 165, 167–176.
  22. Khan, M.A.; Kumar, P.; Akif, M.; Miyoshi, H. Phosphorylation of Eukaryotic Initiation Factor EIFiso4E Enhances the Binding Rates to VPg of Turnip Mosaic Virus. PLoS ONE 2021, 16, e0259688.
  23. Hatta, T.; Bullivant, S.; Matthews, R.E.F. Fine Structure of Vesicles Induced in Chloroplasts of Chinese Cabbage Leaves by Infection with Turnip Yellow Mosaic Virus. J. Gen. Virol. 1973, 20, 37–50.
  24. Kim, K.S. An Ultrastructural Study of Inclusions and Disease Development in Plant Cells Infected by Cowpea Chlorotic Mottle Virus. J. Gen. Virol. 1977, 35, 535–543.
  25. McCartney, A.W.; Greenwood, J.S.; Fabian, M.R.; White, K.A.; Mullen, R.T. Localization of the Tomato Bushy Stunt Virus Replication Protein P33 Reveals a Peroxisome-to-Endoplasmic Reticulum Sorting Pathway. Plant Cell 2005, 17, 3513–3531.
  26. Mochizuki, T.; Hirai, K.; Kanda, A.; Ohnishi, J.; Ohki, T.; Tsuda, S. Induction of Necrosis via Mitochondrial Targeting of Melon Necrotic Spot Virus Replication Protein P29 by Its Second Transmembrane Domain. Virology 2009, 390, 239–249.
  27. Schaad, M.C.; Jensen, P.E.; Carrington, J.C. Formation of Plant RNA Virus Replication Complexes on Membranes: Role of an Endoplasmic Reticulum-Targeted Viral Protein. EMBO J. 1997, 16, 4049–4059.
  28. Amari, K.; Lerich, A.; Schmitt-Keichinger, C.; Dolja, V.V.; Ritzenthaler, C. Tubule-Guided Cell-to-Cell Movement of a Plant Virus Requires Class XI Myosin Motors. PLoS Pathog. 2011, 7, e1002327.
  29. Katsafanas, G.C.; Moss, B. Colocalization of Transcription and Translation within Cytoplasmic Poxvirus Factories Coordinates Viral Expression and Subjugates Host Functions. Cell Host Microbe 2007, 2, 221–228.
  30. Szittya, G.; Molnár, A.; Silhavy, D.; Hornyik, C.; Burgyán, J. Short Defective Interfering RNAs of Tombusviruses Are Not Targeted but Trigger Post-Transcriptional Gene Silencing against Their Helper Virus. Plant Cell 2002, 14, 359–372.
  31. Chellappan, P.; Vanitharani, R.; Ogbe, F.; Fauquet, C.M. Effect of Temperature on Geminivirus-Induced RNA Silencing in Plants. Plant Physiol. 2005, 138, 1828–1841.
  32. Covey, S.N.; Al-Kaff, N.S.; Lángara, A.; Turner, D.S. Plants Combat Infection by Gene Silencing. Nature 1997, 385, 781–782.
  33. Wingard, S.A. Hosts and Symptoms of Ring Spot, A Virus Disease of Plants. J. Agric. Res. 1928, 37, 127–153.
  34. Xin, H.W.; Ding, S.W. Identification and Molecular Characterization of a Naturally Occurring RNA Virus Mutant Defective in the Initiation of Host Recovery. Virology 2003, 317, 253–262.
  35. Ghoshal, B.; Sanfaçon, H. Symptom Recovery in Virus-Infected Plants: Revisiting the Role of RNA Silencing Mechanisms. Virology 2015, 479–480, 167–179.
  36. Ma, X.; Nicole, M.C.; Meteignier, L.V.; Hong, N.; Wang, G.; Moffett, P. Different Roles for RNA Silencing and RNA Processing Components in Virus Recovery and Virus-Induced Gene Silencing in Plants. J. Exp. Bot. 2015, 66, 919–932.
  37. Jovel, J.; Walker, M.; Sanfaçon, H. Recovery of Nicotiana benthamiana Plants from a Necrotic Response Induced by a Nepovirus Is Associated with RNA Silencing but Not with Reduced Virus Titer. J. Virol. 2007, 81, 12285–12297.
  38. Karran, R.A.; Sanfačon, H. Tomato Ringspot Virus Coat Protein Binds to ARGONAUTE 1 and Suppresses the Translation Repression of a Reporter Gene. Mol. Plant-Microbe Interact. 2014, 27, 933–943.
  39. Ratcliff, F.G.; MacFarlane, S.A.; Baulcombe, D.C. Gene Silencing without DNA: RNA-Mediated Cross-Protection between Viruses. Plant Cell 1999, 11, 1207–1215.
  40. Ciomperlik, J.J.; Omarov, R.T.; Scholthof, H.B. An Antiviral RISC Isolated from Tobacco Rattle Virus-Infected Plants. Virology 2011, 412, 117–124.
  41. Kørner, C.J.; Pitzalis, N.; Peña, E.J.; Erhardt, M.; Vazquez, F.; Heinlein, M. Crosstalk between PTGS and TGS Pathways in Natural Antiviral Immunity and Disease Recovery. Nat. Plants 2018, 4, 157–164.
  42. Bhattacharjee, S.; Zamora, A.; Azhar, M.T.; Sacco, M.A.; Lambert, L.H.; Moffett, P. Virus Resistance Induced by NB–LRR Proteins Involves Argonaute4-Dependent Translational Control. Plant J. 2009, 58, 940–951.
  43. Meteignier, L.V.; Zhou, J.; Cohen, M.; Bhattacharjee, S.; Brosseau, C.; Chan, M.G.C.; Robatzek, S.; Moffett, P. NB-LRR Signaling Induces Translational Repression of Viral Transcripts and the Formation of RNA Processing Bodies through Mechanisms Differing from Those Activated by UV Stress and RNAi. J. Exp. Bot. 2016, 67, 2353–2366.
  44. Kohm, B.A.; Goulden, M.G.; Gilbert, J.E.; Kavanagh, T.A.; Baulcombea, D.C. A Potato Virus X Resistance Gene Mediates an Induced, Nonspecific Resistance in Protoplasts. Plant Cell 1993, 5, 913–920.
  45. Richard, M.M.S.; Knip, M.; Schachtschabel, J.; Beijaert, M.S.; Takken, F.L.W. Perturbation of Nuclear–Cytosolic Shuttling of Rx1 Compromises Extreme Resistance and Translational Arrest of Potato Virus X Transcripts. Plant J. 2021, 106, 468–479.
  46. Adams, S.E.; Jones, R.A.C.; Coutts, R.H.A. Expression of Potato Virus X Resistance Gene Rx in Potato Leaf Protoplasts. J. Gen. Virol. 1986, 67, 2341–2345.
  47. Bendahmane, A.; Köhm, B.A.; Dedi, C.; Baulcombe, D.C. The Coat Protein of Potato Virus X Is a Strain-Specific Elicitor of Rx1-Mediated Virus Resistance in Potato. Plant J. 1995, 8, 933–941.
  48. Ritter, E.; Debener, T.; Barone, A.; Salamini, F.; Gebhardt, C. RFLP Mapping on Potato Chromosomes of Two Genes Controlling Extreme Resistance to Potato Virus X (PVX). Mol. Gen. Genet. 1991, 227, 81–85.
  49. Fenyk, S.; Townsend, P.D.; Dixon, C.H.; Spies, G.B.; Campillo, A.D.S.E.; Slootweg, E.J.; Westerhof, L.B.; Gawehns, F.K.K.; Knight, M.R.; Sharples, G.J.; et al. The Potato Nucleotide-Binding Leucine-Rich Repeat (NLR) Immune Receptor Rx1 Is a Pathogen-Dependent DNA-Deforming Protein. J. Biol. Chem. 2015, 290, 24945–24960.
  50. Goulden, M.G.; Baulcombe, D.C. Functionally Homologous Host Components Recognize Potato Virus X in Gomphrena globosa and Potato. Plant Cell 1993, 5, 921–930.
  51. Knip, M.; Richard, M.M.S.; Oskam, L.; van Engelen, H.T.D.; Aalders, T.; Takken, F.L.W. Activation of Immune Receptor Rx1 Triggers Distinct Immune Responses Culminating in Cell Death after 4 Hours. Mol. Plant Pathol. 2019, 20, 575–588.
  52. Slootweg, E.; Roosien, J.; Spiridon, L.N.; Petrescu, A.J.; Tameling, W.; Joosten, M.; Pomp, R.; van Schaik, C.; Dees, R.; Borst, J.W.; et al. Nucleocytoplasmic Distribution Is Required for Activation of Resistance by the Potato NB-LRR Receptor Rx1 and Is Balanced by Its Functional Domains. Plant Cell 2012, 22, 4195–4215.
  53. Sukarta, O.C.A.; Diaz-Granados, A.; Slootweg, E.J.; Overmars, H.; van Schaik, C.; Pokhare, S.; Roosien, J.; Pomp, R.; Elashry, A.; Smant, G.; et al. Two Evolutionary Distinct Effectors from a Nematode and Virus Target RanGAP1 and 2 via the WPP Domain to Promote Disease. bioRxiv 2021.
  54. Hao, W.; Collier, S.M.; Moffett, P.; Chai, J. Structural Basis for the Interaction between the Potato Virus X Resistance Protein (Rx) and Its Cofactor Ran GTPase-Activating Protein 2 (RanGAP2). J. Biol. Chem. 2013, 288, 35868–35876.
  55. Sacco, M.A.; Mansoor, S.; Moffett, P. A RanGAP Protein Physically Interacts with the NB-LRR Protein Rx, and Is Required for Rx-Mediated Viral Resistance. Plant J. 2007, 52, 82–93.
  56. Tameling, W.I.L.; Baulcombe, D.C. Physical Association of the NB-LRR Resistance Protein Rx with a Ran GTPase–Activating Protein Is Required for Extreme Resistance to Potato Virus X. Plant Cell 2007, 19, 1682–1694.
  57. Townsend, P.D.; Dixon, C.H.; Slootweg, E.J.; Sukarta, O.C.A.; Yang, A.W.H.; Hughes, T.R.; Sharples, G.J.; Pålsson, L.O.; Takken, F.L.W.; Goverse, A.; et al. The Intracellular Immune Receptor Rx1 Regulates the DNA-Binding Activity of a Golden2-like Transcription Factor. J. Biol. Chem. 2018, 293, 3218–3233.
  58. Sukarta, O.C.A.; Townsend, P.D.; Llewelyn, A.; Dixon, C.H.; Slootweg, E.J.; Pålsson, L.O.; Takken, F.L.W.; Goverse, A.; Cann, M.J. A DNA-Binding Bromodomain-Containing Protein Interacts with and Reduces Rx1-Mediated Immune Response to Potato Virus X. Plant Commun. 2020, 1, 100086.
  59. Whitham, S.; Dinesh-Kumar, S.P.; Choi, D.; Hehl, R.; Corr, C.; Baker, B. The Product of the Tobacco Mosaic Virus Resistance Gene N: Similarity to Toll and the Interleukin-1 Receptor. Cell 1994, 78, 1101–1115.
  60. Erickson, F.L.; Holzberg, S.; Calderon-Urrea, A.; Handley, V.; Axtell, M.; Corr, C.; Baker, B. The Helicase Domain of the TMV Replicase Proteins Induces the N-Mediated Defence Response in Tobacco. Plant J. 1999, 18, 67–75.
  61. Mestre, P.; Baulcombe, D.C. Elicitor-Mediated Oligomerization of the Tobacco N Disease Resistance Protein. Plant Cell 2006, 18, 491–501.
  62. Caplan, J.L.; Mamillapalli, P.; Burch-Smith, T.M.; Czymmek, K.; Dinesh-Kumar, S.P. Chloroplastic Protein NRIP1 Mediates Innate Immune Receptor Recognition of a Viral Effector. Cell 2008, 132, 449–462.
  63. Burch-Smith, T.M.; Schiff, M.; Caplan, J.L.; Tsao, J.; Czymmek, K.; Dinesh-Kumar, S.P. A Novel Role for the TIR Domain in Association with Pathogen-Derived Elicitors. PLoS Biol. 2007, 5, e68.
  64. Teixeira, R.M.; Ferreira, M.A.; Raimundo, G.A.S.; Loriato, V.A.P.; Reis, P.A.B.; Fontes, E.P.B. Virus Perception at the Cell Surface: Revisiting the Roles of Receptor-like Kinases as Viral Pattern Recognition Receptors. Mol. Plant Pathol. 2019, 20, 1196–1202.
  65. Zorzatto, C.; MacHado, J.P.B.; Lopes, K.V.G.; Nascimento, K.J.T.; Pereira, W.A.; Brustolini, O.J.B.; Reis, P.A.B.; Calil, I.P.; Deguchi, M.; Sachetto-Martins, G.; et al. NIK1-Mediated Translation Suppression Functions as a Plant Antiviral Immunity Mechanism. Nature 2015, 520, 679–682.
  66. Sakamoto, T.; Deguchi, M.; Brustolini, O.J.B.; Santos, A.A.; Silva, F.F.; Fontes, E.P.B. The Tomato RLK Superfamily: Phylogeny and Functional Predictions about the Role of the LRRII-RLK Subfamily in Antiviral Defense. BMC Plant Biol. 2012, 12, 229.
  67. Shan, L.; He, P.; Li, J.; Heese, A.; Peck, S.C.; Nürnberger, T.; Martin, G.B.; Sheen, J. Bacterial Effectors Target the Common Signaling Partner BAK1 to Disrupt Multiple MAMP Receptor-Signaling Complexes and Impede Plant Immunity. Cell Host Microbe 2008, 4, 17–27.
  68. Shiu, S.H.; Bleecker, A.B. Receptor-like Kinases from Arabidopsis Form a Monophyletic Gene Family Related to Animal Receptor Kinases. Proc. Natl. Acad. Sci. USA 2001, 98, 10763–10768.
  69. Carvalho, C.M.; Santos, A.A.; Pires, S.R.; Rocha, C.S.; Saraiva, D.I.; Machado, J.P.B.; Mattos, E.C.; Fietto, L.G.; Fontes, E.P.B. Regulated Nuclear Trafficking of RpL10A Mediated by NIK1 Represents a Defense Strategy of Plant Cells against Virus. PLoS Pathog. 2008, 4, e1000247.
  70. Berlanga, J.J.; Ventoso, I.; Harding, H.P.; Deng, J.; Ron, D.; Sonenberg, N.; Carrasco, L.; Haro, C.D. Antiviral Effect of the Mammalian Translation Initiation Factor 2α Kinase GCN2 against RNA Viruses. EMBO J. 2006, 25, 1730–1740.
  71. Heinicke, L.A.; Wong, C.J.; Lary, J.; Nallagatla, S.R.; Diegelman-Parente, A.; Zheng, X.; Cole, J.L.; Bevilacqua, P.C. RNA Dimerization Promotes PKR Dimerization and Activation. J. Mol. Biol. 2009, 390, 319–338.
  72. Nallagatla, S.R.; Hwang, J.; Toroney, R.; Zheng, X.; Cameron, C.E.; Bevilacqua, P.C. 5′-Triphosphate-Dependent Activation of PKR by RNAs with Short Stem-Loops. Science 2007, 318, 1455–1458.
  73. Toroney, R.; Nallagatla, S.R.; Boyer, J.A.; Cameron, C.E.; Bevilacqua, P.C. Regulation of PKR by HCV IRES RNA: Importance of Domain II and NS5A. J. Mol. Biol. 2010, 400, 393–412.
  74. Brodersen, P.; Sakvarelidze-Achard, L.; Bruun-Rasmussen, M.; Dunoyer, P.; Yamamoto, Y.Y.; Sieburth, L.; Voinnet, O. Widespread Translational Inhibition by Plant MiRNAs and SiRNAs. Science 2008, 320, 1185–1190.
  75. Qu, F.; Morris, T.J. Cap-Independent Translational Enhancement of Turnip Crinkle Virus Genomic and Subgenomic RNAs. J. Virol. 2000, 74, 1085–1093.
  76. Iwakawa, H.O.; Tomari, Y. Molecular Insights into MicroRNA-Mediated Translational Repression in Plants. Mol. Cell 2013, 52, 591–601.
  77. Lanet, E.; Delannoy, E.; Sormani, R.; Floris, M.; Brodersen, P.; Crété, P.; Voinnet, O.; Robaglia, C. Biochemical Evidence for Translational Repression by Arabidopsis MicroRNAs. Plant Cell 2009, 21, 1762–1768.
  78. Dolgin, E. What Lava Lamps and Vinaigrette Can Teach Us about Cell Biology. Nature 2018, 555, 300–302.
  79. Hyman, A.A.; Weber, C.A.; Jülicher, F. Liquid-Liquid Phase Separation in Biology. Annu. Rev. Cell Dev. Biol. 2014, 30, 39–58.
  80. Buchan, J.R.; Parker, R. Eukaryotic Stress Granules: The Ins and Outs of Translation. Mol. Cell 2009, 36, 932–941.
  81. Xu, M.; Mazur, M.J.; Tao, X.; Kormelink, R. Cellular RNA Hubs: Friends and Foes of Plant Viruses. Mol. Plant. Microbe Interact. 2020, 33, 40–54.
  82. Maldonado-Bonilla, L.D. Composition and Function of P Bodies in Arabidopsis thaliana. Front. Plant Sci. 2014, 5, 201.
  83. Brangwynne, C.P.; Eckmann, C.R.; Courson, D.S.; Rybarska, A.; Hoege, C.; Gharakhani, J.; Jülicher, F.; Hyman, A.A. Germline P Granules Are Liquid Droplets That Localize by Controlled Dissolution/Condensation. Science 2009, 324, 1729–1732.
  84. Kroschwald, S.; Munder, M.C.; Maharana, S.; Franzmann, T.M.; Richter, D.; Ruer, M.; Hyman, A.A.; Alberti, S. Different Material States of Pub1 Condensates Define Distinct Modes of Stress Adaptation and Recovery. Cell Rep. 2018, 23, 3327–3339.
  85. Eulalio, A.; Behm-Ansmant, I.; Schweizer, D.; Izaurralde, E. P-Body Formation Is a Consequence, Not the Cause, of RNA-Mediated Gene Silencing. Mol. Cell. Biol. 2007, 27, 3970–3981.
  86. Weber, C.; Nover, L.; Fauth, M. Plant Stress Granules and MRNA Processing Bodies Are Distinct from Heat Stress Granules. Plant J. 2008, 56, 517–530.
  87. Xu, J.; Yang, J.Y.; Niu, Q.W.; Chua, N.H. Arabidopsis DCP2, DCP1, and VARICOSE Form a Decapping Complex Required for Postembryonic Development. Plant Cell 2006, 18, 3386–3398.
  88. Teixeira, D.; Sheth, U.; Valencia-Sanchez, M.A.; Brengues, M.; Parker, R. Processing Bodies Require RNA for Assembly and Contain Nontranslating MRNAs. RNA 2005, 11, 371–382.
  89. Sheth, U.; Parker, R. Decapping and Decay of Messenger RNA Occur in Cytoplasmic Processing Bodies. Science 2003, 300, 805–808.
  90. Beckham, C.J.; Light, H.R.; Nissan, T.A.; Ahlquist, P.; Parker, R.; Noueiry, A. Interactions between Brome Mosaic Virus RNAs and Cytoplasmic Processing Bodies. J. Virol. 2007, 81, 9759–9768.
  91. Galão, R.P.; Chari, A.; Alves-Rodrigues, I.; Lobão, D.; Mas, A.; Kambach, C.; Fischer, U.; Díez, J. LSm1-7 Complexes Bind to Specific Sites in Viral RNA Genomes and Regulate Their Translation and Replication. RNA 2010, 16, 817–827.
  92. Motomura, K.; Le, Q.T.N.; Hamada, T.; Kutsuna, N.; Mano, S.; Nishimura, M.; Watanabe, Y. Diffuse Decapping Enzyme DCP2 Accumulates in DCP1 Foci Under Heat Stress in Arabidopsis thaliana. Plant Cell Physiol. 2015, 56, 107–115.
  93. Lageix, S.; Lanet, E.; Pouch-Pélissier, M.N.; Espagnol, M.C.; Robaglia, C.; Deragon, J.M.; Pélissier, T. Arabidopsis EIF2α Kinase GCN2 Is Essential for Growth in Stress Conditions and Is Activated by Wounding. BMC Plant Biol. 2008, 8, 134.
  94. Carbonell, A.; Carrington, J.C. Antiviral Roles of Plant ARGONAUTES. Curr. Opin. Plant Biol. 2015, 27, 111–117.
  95. Harvey, J.J.W.; Lewsey, M.G.; Patel, K.; Westwood, J.; Heimstädt, S.; Carr, J.P.; Baulcombe, D.C. An Antiviral Defense Role of AGO2 in Plants. PLoS ONE 2011, 6, e14639.
  96. Jaubert, M.; Bhattacharjee, S.; Mello, A.F.S.; Perry, K.L.; Moffett, P. ARGONAUTE2 Mediates RNA-Silencing Antiviral Defenses against Potato Virus X in Arabidopsis. Plant Physiol. 2011, 156, 1556–1564.
  97. Morel, J.B.; Godon, C.; Mourrain, P.; Béclin, C.; Boutet, S.; Feuerbach, F.; Proux, F.; Vaucheret, H. Fertile Hypomorphic ARGONAUTE (Ago1) Mutants Impaired in Post-Transcriptional Gene Silencing and Virus Resistance. Plant Cell 2002, 14, 629–639.
  98. Scholthof, H.B.; Alvarado, V.Y.; Vega-Arreguin, J.C.; Ciomperlik, J.; Odokonyero, D.; Brosseau, C.; Jaubert, M.; Zamora, A.; Moffett, P. Identification of an ARGONAUTE for Antiviral RNA Silencing in Nicotiana benthamiana. Plant Physiol. 2011, 156, 1548–1555.
  99. Qu, F.; Ye, X.; Morris, T.J. Arabidopsis DRB4, AGO1, AGO7, and RDR6 Participate in a DCL4-Initiated Antiviral RNA Silencing Pathway Negatively Regulated by DCL1. Proc. Natl. Acad. Sci. USA 2008, 105, 14732–14737.
  100. Baumberger, N.; Baulcombe, D.C. Arabidopsis ARGONAUTE1 Is an RNA Slicer That Selectively Recruits MicroRNAs and Short Interfering RNAs. Proc. Natl. Acad. Sci. USA 2005, 102, 11928–11933.
  101. Yang, L.; Wu, G.; Poethig, R.S. Mutations in the GW-Repeat Protein SUO Reveal a Developmental Function for MicroRNA-Mediated Translational Repression in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 315–320.
  102. Song, J.J.; Smith, S.K.; Hannon, G.J.; Joshua-Tor, L. Crystal Structure of Argonaute and Its Implications for RISC Slicer Activity. Science 2004, 305, 1434–1437.
  103. Braun, J.E.; Huntzinger, E.; Fauser, M.; Izaurralde, E. GW182 Proteins Directly Recruit Cytoplasmic Deadenylase Complexes to MiRNA Targets. Mol. Cell 2011, 44, 120–133.
  104. Karlowski, W.M.; Zielezinski, A.; Carrère, J.; Pontier, D.; Lagrange, T.; Cooke, R. Genome-Wide Computational Identification of WG/GW Argonaute-Binding Proteins in Arabidopsis. Nucleic Acids Res. 2010, 38, 4231–4245.
  105. Tritschler, F.; Huntzinger, E.; Izaurralde, E. Role of GW182 Proteins and PABPC1 in the MiRNA Pathway: A Sense of Déjà Vu. Nat. Rev. Mol. Cell Biol. 2010, 11, 379–384.
  106. Poulsen, C.; Vaucheret, H.; Brodersen, P. Lessons on RNA Silencing Mechanisms in Plants from Eukaryotic Argonaute Structures. Plant Cell 2013, 25, 22–37.
  107. Kryovrysanaki, N.; James, A.; Tselika, M.; Bardani, E.; Kalantidis, K. RNA silencing pathway in plant development and defense. Int. J. Dev. Biol. 2022, 66, 163–175.
  108. Várallyay, É.; Válóczi, A.; Ágyi, Á.; Burgyán, J.; Havelda, Z. Plant Virus-Mediated Induction of MiR168 Is Associated with Repression of ARGONAUTE1 Accumulation. EMBO J. 2010, 29, 3507–3519.
  109. Azevedo, J.; Garcia, D.; Pontier, D.; Ohnesorge, S.; Yu, A.; Garcia, S.; Braun, L.; Bergdoll, M.; Hakimi, M.A.; Lagrange, T.; et al. Argonaute Quenching and Global Changes in Dicer Homeostasis Caused by a Pathogen-Encoded GW Repeat Protein. Genes Dev. 2010, 24, 904–915.
  110. Zhang, X.; Zhang, X.; Singh, J.; Li, D.; Qu, F. Temperature-Dependent Survival of Turnip Crinkle Virus-Infected Arabidopsis Plants Relies on an RNA Silencing-Based Defense That Requires DCL2, AGO2, and HEN1. J. Virol. 2012, 86, 6847–6854.
  111. Baumberger, N.; Tsai, C.H.; Lie, M.; Havecker, E.; Baulcombe, D.C.C. The Polerovirus Silencing Suppressor P0 Targets ARGONAUTE Proteins for Degradation. Curr. Biol. 2007, 17, 1609–1614.
  112. Iki, T.; Tschopp, M.A.; Voinnet, O. Biochemical and Genetic Functional Dissection of the P38 Viral Suppressor of RNA Silencing. RNA 2017, 23, 639–654.
  113. Thomas, C.L.; Leh, V.; Lederer, C.; Maule, A.J. Turnip Crinkle Virus Coat Protein Mediates Suppression of RNA Silencing in Nicotiana benthamiana. Virology 2003, 306, 33–41.
  114. Schott, G.; Mari-Ordonez, A.; Himber, C.; Alioua, A.; Voinnet, O.; Dunoyer, P. Differential Effects of Viral Silencing Suppressors on SiRNA and MiRNA Loading Support the Existence of Two Distinct Cellular Pools of ARGONAUTE1. EMBO J. 2012, 31, 2553–2565.
  115. Siddiqui, S.A.; Sarmiento, C.; Kiisma, M.; Koivumäki, S.; Lemmetty, A.; Truve, E.; Lehto, K. Effects of Viral Silencing Suppressors on Tobacco Ringspot Virus Infection in Two Nicotiana Species. J. Gen. Virol. 2008, 89, 1502–1508.
  116. Chu, M.; Desvoyes, B.; Turina, M.; Noad, R.; Scholthof, H.B. Genetic Dissection of Tomato Bushy Stunt Virus P19-Protein-Mediated Host-Dependent Symptom Induction and Systemic Invasion. Virology 2000, 266, 79–87.
  117. Lewsey, M.; Surette, M.; Robertson, F.C.; Ziebell, H.; Choi, S.H.; Ryu, K.H.; Canto, T.; Palukaitis, P.; Payne, T.; Walsh, J.A.; et al. The Role of the Cucumber Mosaic Virus 2b Protein in Viral Movement and Symptom Induction. Mol. Plant Pathol. 2009, 22, 642–654.
  118. Lin, S.S.; Wu, H.W.; Jan, F.J.; Hou, R.F.; Yeh, S.D. Modifications of the Helper Component-Protease of Zucchini Yellow Mosaic Virus for Generation of Attenuated Mutants for Cross Protection Against Severe Infection. Phytopathology 2007, 97, 287–296.
  119. Wu, Q.; Wang, X.; Ding, S.-W. Viral Suppressors of RNA-Based Viral Immunity: Host Targets. Cell Host Microbe 2010, 8, 12–15.
  120. Fernández-Calvino, L.; Martínez-Priego, L.; Szabo, E.Z.; Guzmán-Benito, I.; González, I.; Canto, T.; Lakatos, L.; Llave, C. Tobacco Rattle Virus 16K Silencing Suppressor Binds ARGONAUTE 4 and Inhibits Formation of RNA Silencing Complexes. J. Gen. Virol. 2016, 97, 246–257.
  121. Chan, S.W.L.; Zilberman, D.; Xie, Z.; Johansen, L.K.; Carrington, J.C.; Jacobsen, S.E. RNA Silencing Genes Control de Novo DNA Methylation. Science 2004, 303, 1336.
  122. Zilberman, D.; Cao, X.; Jacobsen, S.E. ARGONAUTE4 Control of Locus-Specific SiRNA Accumulation and DNA and Histone Methylation. Science 2003, 299, 716–719.
  123. Brosseau, C.; Oirdi, M.E.; Adurogbangba, A.; Ma, X.; Moffett, P. Antiviral Defense Involves AGO4 in an Arabidopsis-Potexvirus Interaction. Mol. Plant. Microbe Interact. 2016, 29, 878–888.
  124. Silhavy, D.; Molnár, A.; Lucioli, A.; Szittya, G.; Hornyik, C.; Tavazza, M.; Burgyán, J. A Viral Protein Suppresses RNA Silencing and Binds Silencing-Generated, 21- to 25-Nucleotide Double-Stranded RNAs. EMBO J. 2002, 21, 3070–3080.
More
Information
Subjects: Plant Sciences
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 292
Revisions: 2 times (View History)
Update Date: 07 Jul 2023
1000/1000
ScholarVision Creations