#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Molecular characterization and In Vitro synthesis of infectious RNA of a Turnip vein-clearing virus isolated from Alliaria petiolata in Hungary


Authors: Tamás Tóth aff001;  Péter Gyula aff001;  Pál Salamon aff002;  Szilvia Kis aff001;  Anita Sós-Hegedűs aff001;  György Szittya aff001
Authors place of work: Department of Plant Biotechnology, Agricultural Biotechnology Institute, National Agricultural Research and Innovation Center, Gödöllő, Hungary aff001;  Department of Genetics, Agricultural Biotechnology Institute, National Agricultural Research and Innovation Center, Gödöllő, Hungary aff002
Published in the journal: PLoS ONE 14(10)
Category: Research Article
doi: https://doi.org/10.1371/journal.pone.0224398

Summary

A tobamovirus was isolated from leaves of Alliaria petiolata plants, showing vein-clearing, interveinal chlorosis, and moderate deformation. Host range experiments revealed a high similarity of isolate ApH both to ribgrass mosaic viruses and turnip vein-clearing viruses. The complete nucleotide sequence of the viral genome was determined. The genomic RNA is composed of 6312 nucleotides and contains four open reading frames (ORF). ORF1 is 3324 nt-long and encodes a polypeptide of about 125.3 kDa. The ORF1 encoded putative replication protein contains an Alphavirus-like methyltransferase domain. ORF2 is 4806 nt-long and encodes a polypeptide of about 182 kDa. The ORF2 encoded putative replication protein contains an RNA-dependent RNA polymerase, catalytic domain. ORF3 encodes the putative cell-to-cell movement protein with a molecular weight of 30.1 kDa. ORF4 overlaps with ORF3 and encodes the coat protein with a size of 17.5 kDa. Sequence comparisons revealed that the ApH isolate has the highest similarity to turnip vein-clearing viruses and should be considered an isolate of Turnip vein-clearing virus (TVCV). This is the first report on the occurrence of TVCV in Hungary. In vitro transcripts prepared from the full-length cDNA clone of TVCV-ApH were highly infectious and induced typical symptoms characteristic to the original isolate of the virus. Since infectious clones of TVCV-ApH and crTMV (another isolate of TVCV) markedly differed in respect to recovery phenotype in Arabidopsis thaliana, it is feasible to carry out gene exchange or mutational studies to determine viral factors responsible for the symptom recovery phenotype.

Keywords:

Plant genomics – Phylogenetic analysis – RNA viruses – Viral replication – Viral genomics – Arabidopsis thaliana – Nicotiana – Plant viral pathogens

Introduction

The genus Tobamovirus (family Virgaviridae) represents a group of rod-shaped (approx. 300–310 nm long) plant viruses with single-stranded, undivided positive-sense RNA genomes of approximately 6400 nucleotides. There are currently 37 species in this genus including the type species Tobacco mosaic virus (TMV) which is one of the best-studied and characterized plant viruses [1]. The genome of tobamoviruses contains at least four ORFs. ORF1 codes for the 124–132 kDa small replicase subunit, while the suppression of an amber termination codon of ORF1 results in the 181–189 kDa large replicase subunit (ORF2). The ORF3 and ORF4 translate from 3’ subgenomic RNAs and result in a 28–31 kDa movement protein, and a 17–18 kDa coat protein, respectively [2]. Several reports demonstrated that besides its role in replication, the p124–132 protein acts as a viral silencing suppressor through the binding of viral siRNAs [35].

The genus Tobamovirus is divided at least into three subgroups based on natural host range, genomic organization and phylogenetic clustering of isolates [68]. The Subgroup 3 includes crucifer- and plantain-infecting viruses and currently comprises the following four species: Ribgrass mosaic virus (RMV), Turnip vein-clearing virus (TVCV), Youcai mosaic virus (YoMV) formerly designated as Oilseed rape mosaic virus (ORMV) and Wasabi mosaic virus (WMoV) [1]. TVCV was first identified from infected turnip (Brassica rapa L.) plants in Oklahoma [9] as a contaminant of a preparation of cauliflower mosaic virus (CaMV). At the time that TVCV was first being characterized by Lartey et al., (1993) [9,10] a related virus isolate was also being described by Dorokhov et al., [1113]. Dorokhov et al. designated their isolate as crTMV (for crucifer TMV), however, it differs only 6.5% in the nucleotide sequence from the TVCV isolate [14] and therefore it is regarded as an isolate of TVCV [15].

In this work, we determined the complete nucleotide sequence of the genomic RNA of TVCV isolated from Alliaria petiolata in Hungary and report the preparation of the infectious cDNA clone of the virus.

Materials and methods

Collection, host range determination, and maintenance of the virus

The virus was isolated from Alliaria petiolata in Hungary and maintained in Nicotiana benthamiana until an infectious transcript was created. The host range was studied by inoculating the following test plants: Arabidopsis thaliana Col-0 and Bur-0 ecotypes, Brassica chinensis, Brassica rapa var. rapa, Capsicum annuum, Chenopodium murale, Chenopodium quinoa, Datura stramonium, Nicotiana benthamiana, Nicotiana glutinosa, Nicotiana megalosiphon, Nicotiana sylvestris, Nicotiana tabacum cv. Samsun, Nicotiana tabacum cv. Xanthi, Ocimum basilicum, Solanum betaceum, Solanum lycopersicum.

Cloning of the virus and the making of an infectious transcript

Virions and viral genomic RNA were purified from infected Nicotiana benthamiana plants according to a tobamovirus isolation protocol [16]. The viral RNA was converted to cDNA by SuperScript IV reverse transcriptase (Invitrogen) using the virus-specific primer RMV 3 (5’-TGG GCC CCT ACC CGG GGT TAG GGA GG-3’). The viral genome was amplified with CloneAmp HiFi PCR Premix (Clontech) in one piece using the primers RMV Inf Fw (5’-CGG TAC CCG GGG ATC GTT TAG TTT TAT TGC AAC AAC AAC A-3’) and RMV Inf Rev (5’-CGA CTC TAG AGG ATC TGG GCC CCT ACC CGG GGT-3’). The PCR products were cloned into pUC19 vector using the In-Fusion® HD Cloning Kit (Clontech). The cloned viral genome was sequenced by the Sanger method in eight parts and the sequences were assembled by CAP3 [17]. The homology search was performed using the BLAST server [18]. The full-length viral sequence was deposited in GenBank under the accession number MH370485.

The infectious transcript was created in the following way: a PCR was carried out with a primer specific to the 5’ end of the virus-containing a T7 promoter as a 5’ flag (RMV 5’ T7 BamHI, 5’-GAG AGG ATC CTA ATA CGA CTC ACT ATA GGT TTA GTT TTA TTG CAA CAA CAA-3’), and another primer specific to a region in the virus containing a natural StuI recognition site (RMV 632 minus, 5’-ATC AGT ACC GTT CGC CGT CG T-3’). The PCR product was phosphorylated with T4 polynucleotide kinase (Thermo Scientific), digested with StuI, and ligated into the pUC19-TVCV vector that was first KpnI digested, blunted with T4 DNA polymerase, digested with StuI, and then dephosphorylated with Antarctic phosphatase (New England Biolabs). The 3’ end of the virus was edited to include a BstXI restriction enzyme recognition site to allow linearization of the vector for in vitro transcription. For this, the 3’ end of the virus was amplified with TVCV MP fw (5’-CGG GTT GGC AGC CGT TAG CTC TGG-3’) and TVCV 3’ BstXI (5’-TAG CCA ATT ATT TGG GCC CCT ACC CGG GGT TAG GGA GG-3’) primers, the product was phosphorylated with T4 polynucleotide kinase, cut with BstEII, and ligated into the pUC19-T7::TVCV that was cut with SphI, blunted with T4 DNA polymerase, cut with BstEII, and dephosphorylated with Antarctic phosphatase. The positive clones were verified by Sanger sequencing. For the in vitro transcription, the pUC19-T7::TVCV-BstXI vector was linearized with BstXI, and transcribed with T7 RNA polymerase either with or without the 7mG cap analog (New England Biolabs) as described earlier [19].

Inoculation of plants

For every leaf, 4 μl of the in vitro transcription reaction mixture was combined with 3.5 μl water and 7.5 μl 2× inoculation buffer containing 50 mM glycine, 30 mM K₂HPO₄ pH 9.2, 1% (w/v) Bentonite-SF (Serva), 1% (w/v) Celite® 512 (Fluka). The inoculation solution was applied onto the leaves of four-week-old Nicotiana benthamiana plants and rub inoculated with a glass spatula. After inoculation, the plants were sprayed with water. For the inoculation of Arabidopsis thaliana plants, 2 μl (1 μg) of total RNA purified from TVCV-ApH or crTMV infected systemic leaves of N. benthamiana was combined with 5.5 μl water and 7.5 μl 2× inoculation buffer and was rub inoculated into the true leaves of four-week-old Arabidopsis plants. Three leaves were inoculated per plant. After inoculation, the plants were sprayed with water.

Total RNA purification and northern blot analysis

Total RNA from plants was purified using a phenol-chloroform extraction method [20]. A probe recognizing the 3’-end of the virus was amplified with primers TVCV MP fw and TVCV 3’ BstXI (see above). The labeling and the northern blot analysis were carried out according to Baksa et al., 2015 [21].

Protein purification and PAGE

Proteins were purified as described earlier [22] and separated by electrophoresis on 12% TGX Stain-Free FastCast Acrylamide Gels (Bio-Rad) using ProSieve QuadColor protein marker (4.6–300 kDa) (Lonza).

Phylogenetic analysis

The full-length genomes of tobamoviruses were retrieved from the NCBI Reference Sequence Database [23] and aligned with the MUSCLE algorithm [24]. The phylogenetic relationship was inferred using the Neighbor-Joining method [25] as implemented in the MEGA X software [26]. The accuracy of the presented tree structure was estimated by the bootstrap method (1000 replicates) [27]. The trees were drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the LogDet (Tamura-Kumar) method [28] and are in the units of the number of base substitutions per site. The rate variation among sites was modeled with a gamma distribution (shape parameter = 0.05). The differences in the composition bias among sequences were considered in evolutionary comparisons. All positions containing gaps and missing data were eliminated.

Results and discussion

Isolation and host range determination of the virus

The Alliaria petiolata plant showing vein-clearing, chlorotic regions, and moderate leaf deformation (Fig 1) was collected at a roadside in Velence, Hungary. Several test plants inoculated by its leaf extract reacted either with local or with local and systemic symptoms characteristic to tobamoviruses. A single lesion subculture of the isolated virus caused symptoms on selected test plants (Fig 1 and Table 1) very similar to those of RMV, the only tobamovirus known to be pathogenic to cruciferous plants and Solanum betaceum in Hungary.

Symptoms observed in different plants infected by TVCV-ApH.
Fig. 1. Symptoms observed in different plants infected by TVCV-ApH.
(a) Mild vein clearing and mosaic in the leaf of garlic mustard (Alliaria petiolata) the source of TVCV-ApH. (b) Systemic symptoms on Solanum betaceum. (c) Nicotiana tabacum cv. Samsun, systemic symptoms. (d) Local lesions on Nicotiana tabacum cv. Xanthi-nc.
Tab. 1. Host range and symptomatology of TVCV-ApH.
Host range and symptomatology of TVCV-ApH.

Cloning and sequencing of the virus

To further characterize the ApH isolate, virions were purified from systemically infected N. benthamiana leaves [16]. RNA extraction and genome amplification by RT-PCR, cloning, sequencing, the assembly of the full genome sequence and gene translations of the ApH isolate were conducted as described in the materials and methods section. A sequence homology search with the full-length genomic sequence revealed that the closest relatives of the ApH isolate are various isolates of Turnip vein-clearing virus and Ribgrass mosaic virus.

Phylogenetic relationship of the ApH isolate and other members of the Tobamovirus family

Phylogenetic analysis of the full-length tobamovirus genomes from the RefSeq database confirmed that the ApH isolate is most closely related to TVCV and belongs to Subgroup 3 of the genus Tobamovirus (Fig 2). Based on the molecular and phylogenetic data we conclude that isolate ApH is a strain of TVCV and propose to name it TVCV-ApH.

Phylogenetic relationship of the TVCV-ApH and some representative members of tobamoviruses.
Fig. 2. Phylogenetic relationship of the TVCV-ApH and some representative members of tobamoviruses.
The evolutionary history was inferred using the Neighbor-Joining method [25]. The optimal tree with the sum of branch length = 8.01862305 is shown. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) is shown next to the branches [27]. All positions containing gaps and missing data were eliminated. There were a total of 5052 positions in the final dataset. Tobacco rattle virus was used as an outgroup.

Comparison of TVCV isolates with full genomes

Currently, four complete TVCV genome sequences are available in the GenBank database (NC_001873.1, Z29370, JN205074.1, JN205073.1). Based on phylogenetic analyses using full genome nucleotide sequences, the closest relative of TVCV-ApH is the isolate that was formerly known as crTMV [12] (Fig 3).

Phylogenetic relationship of the TVCV-ApH and the TVCV isolates with available full genome sequences.
Fig. 3. Phylogenetic relationship of the TVCV-ApH and the TVCV isolates with available full genome sequences.
The evolutionary history was inferred using the Neighbor-Joining method [25]. The optimal tree with the sum of branch length = 0.16283470 is shown. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches [27]. All positions containing gaps and missing data were eliminated. There were a total of 5965 positions in the final dataset. Ribgrass mosaic virus was used as an outgroup.

Comparison of the viral-encoded proteins of TVCV-ApH and the four TVCV isolates revealed a moderate amount of differences in the amino acid sequences (S1 Fig).

Genome organization of TVCV-ApH

The genome of TVCV-ApH is 6312-nt-long containing four open reading frames (ORF) (Fig 4). The first ORF starts from an AUG at nts 68–70 and terminates with an amber stop codon UAG at nts 3389–3391. Readthrough of the leaky amber termination codon would extend the frame up to a stop codon at nt 4871–4873 (ORF2). ORF1 is 3324-nt-long and encodes a polypeptide of about 125.3 kDa. The ORF1 encoded putative replication protein contains an Alphavirus-like methyltransferase (MTR) domain (position 72–280 aa; PS51743) and (+) RNA virus helicase core domain (position 795–1107 aa; PS51657) [29]. ORF2 is 4806-nt-long (68–4873 nt) and encodes a polypeptide of about 182 kDa. The ORF2 encoded putative replication protein contains an RNA-directed RNA polymerase, catalytic domain (position 1368–1481 aa; PS50507) [29]. ORF3 starts from nt 4876 and terminates with a UAA at nts 5677–5679. ORF3 encodes the putative cell-to-cell movement protein (MP) with a size of ca. 30.1 kDa. ORF4 overlaps with ORF3 and starts from nt 5603 and terminates at nt 6076 and encodes the coat protein with a size of 17.5 kDa. The 5’ untranslated region is 67-nt-long and the 3’ untranslated region is 236-nt-long.

Schematic representation of the genome organization of TVCV-ApH.
Fig. 4. Schematic representation of the genome organization of TVCV-ApH.
Similarly to other tobamoviruses, the TVCV-ApH genome contains four ORFs. The genomic RNA is a template for expression of the 125 and 182 kDa replication proteins. The 3’ distal movement protein (MP) and capsid protein (CP) ORFs are expressed from separate 3’ co-terminal sgRNAs. The tRNA structure motif at the 3’-end of the RNA is represented by a cloverleaf motif.

Characterization of the infectious in vitro transcribed viral RNA

Full-length clone of TVCV-ApH was constructed with a T7 promoter in front of the genomic sequence of the virus (pUC19-T7::TVCV-BstXI) as described in the materials and methods section. The full-length capped and uncapped in vitro transcripts of TVCV-ApH were prepared by transcription of BstXI-linearized pUC19-T7::TVCV-BstXI plasmid with T7 RNA polymerase. The in vitro transcribed viral RNAs co-migrated with the TVCV-ApH genomic RNA extracted from virus-infected plants (Fig 5A). In vitro transcripts of both capped and uncapped TVCV-ApH RNAs were infectious; however, infectivity was greatly reduced and virus accumulation delayed in the absence of the cap structure (Fig 5B and 5C). All plants inoculated either with the wild-type viral RNA or the in vitro synthesized, capped TVCV-ApH RNA showed similar symptoms, which developed with the same intensity and rapidity. Furthermore, northern blot analyses confirmed that the patterns of viral-specific genomic and subgenomic RNA species were equivalent to the wild-type virus (Fig 5C). Moreover, SDS-PAGE analyses of the protein extract from in vitro transcribed TVCV-ApH RNA-infected plants confirmed the accumulation of a protein of ca. 18 kDa, which is characteristic of TVCV coat protein (Fig 5D). Therefore we concluded that the in vitro transcribed TVCV-ApH RNA is biologically active and pathologically indistinguishable from the parent virus.

Infection of <i>Nicotiana benthamiana</i> plants with the <i>in vitro</i> transcribed TVCV-ApH.
Fig. 5. Infection of Nicotiana benthamiana plants with the in vitro transcribed TVCV-ApH.
(a) Agarose gel electrophoresis of in vitro synthesized infectious transcripts with or without the m7G cap analog. (b) Denaturing agarose gel electrophoresis of total RNAs from mock-treated, uncapped-, and capped-transcript-treated Nicotiana benthamiana plants after 5 dpi. The gel was stained with ethidium bromide. (c) The same gel was blotted to a membrane and subjected to northern blot analysis using a virus-specific probe. The genomic (gRNA) and the subgenomic RNAs (sgRNA1 and sgRNA2) are marked. (d) Total proteins were purified from the inoculated and the systemic leaves of the same mock- and capped-transcript-treated plants as in the case of northern analysis and run on 12% TGX Stain-Free FastCast Acrylamide Gels (Bio-Rad). The 18 kDa Coat Protein is marked.

Symptom recovery can be observed in plants infected with crTMV but not with TVCV-ApH

ORMV (also called Youcai mosaic virus or Chinese rape mosaic virus, or TMV-Cg) like TVCV, belongs to the Subgroup 3 of the tobamoviruses that causes symptoms in crucifers and is a commonly used virus model for studies in Arabidopsis thaliana [30]. It was reported recently, that A. thaliana plants infected with ORMV undergo natural symptom recovery which is characterized by the emergence of asymptomatic leaves following a systemic symptomatic infection [31]. To test whether TVCV-ApH infection also results in symptom recovery, we inoculated A. thaliana ecotype Bur-0 plants either with in vitro transcripts of TVCV-ApH or crTMV [12], the closest relative of TVCV-ApH, according to our phylogenetic analysis (Fig 3). TVCV-ApH has a 96.15% nucleotide identity (242 SNPs) compared to the crTMV at the full genome level. Multiple alignment of the viral-encoded proteins of the TVCV isolates with full genomes including TVCV-ApH and crTMV is in S1 Fig.

Both TVCV-ApH- and crTMV-infected plants showed disease symptoms at 14 days post-inoculation (dpi) and equally high level of virus accumulation in the systemically infected leaves by northern blot analysis (Fig 6). Similarly to the ORMV-infected plants reported previously [31], from the fourth week after inoculation, all newly emerging leaves recovered in 9 crTMV-infected plants out of the 10 plants tested, while none of the TVCV-ApH-infected plants recovered. The virus level also remained high in the symptomatic leaves, while decreased in the symptomless ones (Fig 6).

Infection of <i>Arabidopsis thaliana</i> Bur-0 plants with TVCV-ApH- and crTMV.
Fig. 6. Infection of Arabidopsis thaliana Bur-0 plants with TVCV-ApH- and crTMV.
(a) Symptom development and recovery in the Bur-0 ecotype of Arabidopsis thaliana plants infected either with TVCV-ApH or crTMV. (b) The viral genomic (gRNA) and subgenomic RNA (sgRNA1 and sgRNA2) levels were detected in a northern blot assay using a probe that recognizes the 3’ end of the viruses which is highly conserved at the nucleotide level. The loading was checked by running the total RNAs on an agarose gel that was stained with ethidium bromide. For quantitative analysis, the density of the gRNA northern signal was measured in every sample and was normalized to the 28S ribosomal RNA (28S rRNA) levels in the corresponding sample. The values were further normalized to the first sample. The resulting relative values were placed under the samples.

The infectious clone of TVCV-ApH may serve the scientific community

The available infectious cDNA clone of TVCV-ApH can be useful for further studies to identify viral factors responsible for the marked differences in symptom recovery phenotype between TVCV-ApH and the closely related crTMV.

Supporting information

S1 Fig [pdf]
Multiple alignment of the virus-encoded proteins in TVCV isolates.

S1 Raw Images [pdf]
Original, uncropped images of gels and blots presented in figures.


Zdroje

1. Lefkowitz EJ, Dempsey DM, Hendrickson RC, Orton RJ, Siddell SG, Smith DB. Virus taxonomy: the database of the International Committee on Taxonomy of Viruses (ICTV). Nucleic Acids Res. 2018;46: D708–D717. doi: 10.1093/nar/gkx932 29040670

2. Ishibashi K, Ishikawa M. Replication of Tobamovirus RNA. Annu Rev Phytopathol. 2016;54: 55–78. doi: 10.1146/annurev-phyto-080615-100217 27296148

3. Csorba T, Bovi A, Dalmay T, Burgyán J. The p122 subunit of Tobacco mosaic virus replicase is a potent silencing suppressor and compromises both small interfering RNA- and microRNA-mediated pathways. J Virol. 2007;81: 11768–11780. doi: 10.1128/JVI.01230-07 17715232

4. Vogler H, Akbergenov R, Shivaprasad PV, Dang V, Fasler M, Kwon M-O, et al. Modification of small RNAs associated with suppression of RNA silencing by tobamovirus replicase protein. J Virol. 2007;81: 10379–10388. doi: 10.1128/JVI.00727-07 17634237

5. Ding XS, Liu J, Cheng N-H, Folimonov A, Hou Y-M, Bao Y, et al. The Tobacco mosaic virus 126-kDa protein associated with virus replication and movement suppresses RNA silencing. Mol Plant-Microbe Interact MPMI. 2004;17: 583–592. doi: 10.1094/MPMI.2004.17.6.583 15195941

6. Chavan RR, Pearson MN. Molecular characterisation of a novel recombinant Ribgrass mosaic virus strain FSHS. Virol J. 2016;13: 29. doi: 10.1186/s12985-016-0487-5 26891841

7. Gibbs AJ, Wood J, Garcia-Arenal F, Ohshima K, Armstrong JS. Tobamoviruses have probably co-diverged with their eudicotyledonous hosts for at least 110 million years. Virus Evol. 2015;1: vev019. doi: 10.1093/ve/vev019 27774289

8. Heinze C, Lesemann D-E, Ilmberger N, Willingmann P, Adam G. The phylogenetic structure of the cluster of tobamovirus species serologically related to ribgrass mosaic virus (RMV) and the sequence of streptocarpus flower break virus (SFBV). Arch Virol. 2006;151: 763–774. doi: 10.1007/s00705-005-0640-8 16328151

9. Lartey RT. Occurrence of a vein-clearing tobamovirus in turnip. Plant Dis. 1993;77: 21. doi: 10.1094/PD-77-0021

10. Lartey RT, Lane LC, Melcher U. Electron microscopic and molecular characterization of turnip vein-clearing virus. Arch Virol. 1994;138: 287–298. doi: 10.1007/bf01379132 7998835

11. Dorokhov IL, Ivanov PA, Novikov VK, Yefimov VA, Atabekov IG. [Tobamovirus of cruciferous plants: nucleotide sequence of genes of the transport protein, capsid protein, and 3’-terminal untranslated region]. Dokl Akad Nauk. 1993;332: 518–522. 8260923

12. Dorokhov YL, Ivanov PA, Novikov VK, Agranovsky AA, Morozov SY, Efimov VA, et al. Complete nucleotide sequence and genome organization of a tobamovirus infecting cruciferae plants. FEBS Lett. 1994;350: 5–8. doi: 10.1016/0014-5793(94)00721-7 7545946

13. Dorokhov IL, Ivanov PA, Novikov VK, Efimov VA, Atabekov IG. [Tobamovirus of the Cruciferae family: nucleotide sequence of the 5’-untranslated region and nonstructural protein genes controlling replication viral genome]. Dokl Akad Nauk. 1994;335: 792–798. 8025551

14. Lartey RT, Voss TC, Melcher U. Completion of a cDNA sequence from a tobamovirus pathogenic to crucifers. Gene. 1995;166: 331–332. doi: 10.1016/0378-1119(95)00674-5 8543186

15. Melcher U. Turnip vein-clearing virus, from pathogen to host expression profile. Mol Plant Pathol. 2003;4: 133–140. doi: 10.1046/j.1364-3703.2003.00159.x 20569373

16. Chapman SN. Tobamovirus isolation and RNA extraction. In: Foster GD, Taylor SC, editors. Plant Virology Protocols: From Virus Isolation to Transgenic Resistance. Totowa, NJ: Humana Press; 1998. pp. 123–129. doi: 10.1385/0-89603-385-6:123

17. Huang X, Madan A. CAP3: A DNA sequence assembly program. Genome Res. 1999;9: 868–877. doi: 10.1101/gr.9.9.868 10508846

18. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215: 403–410. doi: 10.1016/S0022-2836(05)80360-2 2231712

19. 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. doi: 10.1105/tpc.010366 11884680

20. Szittya G, Salamon P, Burgyán J. The complete nucleotide sequence and synthesis of infectious RNA of genomic and defective interfering RNAs of TBSV-P. Virus Res. 2000;69: 131–136. doi: 10.1016/s0168-1702(00)00178-7 11018282

21. Baksa I, Nagy T, Barta E, Havelda Z, Várallyay É, Silhavy D, et al. Identification of Nicotiana benthamiana microRNAs and their targets using high throughput sequencing and degradome analysis. BMC Genomics. 2015;16: 1025. doi: 10.1186/s12864-015-2209-6 26626050

22. Kis S, Salamon P, Kis V, Szittya G. Molecular characterization of a beet ringspot nepovirus isolated from Begonia ricinifolia in Hungary. Arch Virol. 2017;162: 3559–3562. doi: 10.1007/s00705-017-3521-z 28812162

23. Brister JR, Ako-Adjei D, Bao Y, Blinkova O. NCBI viral genomes resource. Nucleic Acids Res. 2015;43: D571–577. doi: 10.1093/nar/gku1207 25428358

24. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32: 1792–1797. doi: 10.1093/nar/gkh340 15034147

25. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4: 406–425. doi: 10.1093/oxfordjournals.molbev.a040454 3447015

26. Kumar S, Stecher G, Li M, Knyaz C, Tamura K, Battistuzzi FU. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol Biol Evol. 2018;35: 1547–1549. doi: 10.1093/molbev/msy096 29722887

27. Felsenstein J. Confidence limits on phylogenies: an approach using the bootstrap. Evol Int J Org Evol. 1985;39: 783–791. doi: 10.1111/j.1558-5646.1985.tb00420.x 28561359

28. Tamura K, Kumar S. Evolutionary distance estimation under heterogeneous substitution pattern among lineages. Mol Biol Evol. 2002;19: 1727–1736. doi: 10.1093/oxfordjournals.molbev.a003995 12270899

29. Mitchell AL, Attwood TK, Babbitt PC, Blum M, Bork P, Bridge A, et al. InterPro in 2019: improving coverage, classification and access to protein sequence annotations. Nucleic Acids Res. 2019;47: D351–D360. doi: 10.1093/nar/gky1100 30398656

30. Lartey RT, Voss TC, Melcher U. Tobamovirus evolution: gene overlaps, recombination, and taxonomic implications. Mol Biol Evol. 1996;13: 1327–1338. doi: 10.1093/oxfordjournals.molbev.a025579 8952077

31. Kørner CJ, Pitzalis N, Peña EJ, 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. doi: 10.1038/s41477-018-0117-x 29497161


Článek vyšel v časopise

PLOS One


2019 Číslo 10
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

Svět praktické medicíny 1/2024 (znalostní test z časopisu)
nový kurz

Koncepce osteologické péče pro gynekology a praktické lékaře
Autoři: MUDr. František Šenk

Sekvenční léčba schizofrenie
Autoři: MUDr. Jana Hořínková

Hypertenze a hypercholesterolémie – synergický efekt léčby
Autoři: prof. MUDr. Hana Rosolová, DrSc.

Význam metforminu pro „udržitelnou“ terapii diabetu
Autoři: prof. MUDr. Milan Kvapil, CSc., MBA

Všechny kurzy
Kurzy Podcasty Doporučená témata Časopisy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#