150 research outputs found

    Gun1 controls accumulation of the plastid ribosomal protein S1 at the protein level and interacts with proteins involved in plastid protein homeostasis

    Get PDF
    Developmental or metabolic changes in chloroplasts can have profound effects on the rest of the plant cell. Such intracellular responses are associated with signals that originate in chloroplasts and convey information on their physiological status to the nucleus, which leads to large-scale changes in gene expression (retrograde signaling). A screen designed to identify components of retrograde signaling resulted in the discovery of the so-called genomes uncoupled (gun) mutants. Genetic evidence suggests that the chloroplast protein GUN1 integrates signals derived from perturbations in plastid redox state, plastid gene expression, and tetrapyrrole biosynthesis (TPB) in Arabidopsis (Arabidopsis thaliana) seedlings, exerting biogenic control of chloroplast functions. However, the molecular mechanism by which GUN1 integrates retrograde signaling in the chloroplast is unclear. Here we show that GUN1 also operates in adult plants, contributing to operational control of chloroplasts. The gun1 mutation genetically interacts with mutations of genes for the chloroplast ribosomal proteins S1 (PRPS1) and L11. Analysis of gun1 prps1 lines indicates that GUN1 controls PRPS1 accumulation at the protein level. The GUN1 protein physically interacts with proteins involved in chloroplast protein homeostasis based on coimmunoprecipitation experiments. Furthermore, yeast two-hybrid and bimolecular fluorescence complementation experiments suggest that GUN1 might transiently interact with several TPB enzymes, including Mg-chelatase subunit D (CHLD) and two other TPB enzymes known to activate retrograde signaling. Moreover, the association of PRPS1 and CHLD with protein complexes is modulated by GUN1. These findings allow us to speculate that retrograde signaling might involve GUN1-dependent formation of protein complexes

    Sense and Antisense Transcripts of Convergent Gene Pairs in Arabidopsis thaliana Can Share a Common Polyadenylation Region

    Get PDF
    The Arabidopsis genome contains a large number of gene pairs that encode sense and antisense transcripts with overlapping 3′ regions, indicative for a potential role of natural antisense transcription in regulating sense gene expression or transcript processing. When we mapped poly(A) transcripts of three plant gene pairs with long overlapping antisense transcripts, we identified an unusual transcript composition for two of the three gene pairs. Both genes pairs encoded a class of long sense transcripts and a class of short sense transcripts that terminate within the same polyadenylation region as the antisense transcripts encoded by the opposite strand. We find that the presence of the short sense transcript was not dependent on the expression of an antisense transcript. This argues against the assumption that the common termination region for sense and antisense poly(A) transcripts is the result of antisense-specific regulation. We speculate that for some genes evolution may have especially favoured alternative polyadenylation events that shorten transcript length for gene pairs with overlapping sense/antisense transcription, if this reduces the likelihood for dsRNA formation and transcript degradation

    Diverse Forms of RPS9 Splicing Are Part of an Evolving Autoregulatory Circuit

    Get PDF
    Ribosomal proteins are essential to life. While the functions of ribosomal protein-encoding genes (RPGs) are highly conserved, the evolution of their regulatory mechanisms is remarkably dynamic. In Saccharomyces cerevisiae, RPGs are unusual in that they are commonly present as two highly similar gene copies and in that they are over-represented among intron-containing genes. To investigate the role of introns in the regulation of RPG expression, we constructed 16 S. cerevisiae strains with precise deletions of RPG introns. We found that several yeast introns function to repress rather than to increase steady-state mRNA levels. Among these, the RPS9A and RPS9B introns were required for cross-regulation of the two paralogous gene copies, which is consistent with the duplication of an autoregulatory circuit. To test for similar intron function in animals, we performed an experimental test and comparative analyses for autoregulation among distantly related animal RPS9 orthologs. Overexpression of an exogenous RpS9 copy in Drosophila melanogaster S2 cells induced alternative splicing and degradation of the endogenous copy by nonsense-mediated decay (NMD). Also, analysis of expressed sequence tag data from distantly related animals, including Homo sapiens and Ciona intestinalis, revealed diverse alternatively-spliced RPS9 isoforms predicted to elicit NMD. We propose that multiple forms of splicing regulation among RPS9 orthologs from various eukaryotes operate analogously to translational repression of the alpha operon by S4, the distant prokaryotic ortholog. Thus, RPS9 orthologs appear to have independently evolved variations on a fundamental autoregulatory circuit

    Parasite-insecticide interactions: a case study of Nosema ceranae and fipronil synergy on honeybee

    Get PDF
    In ecosystems, a variety of biological, chemical and physical stressors may act in combination to induce illness in populations of living organisms. While recent surveys reported that parasite-insecticide interactions can synergistically and negatively affect honeybee survival, the importance of sequence in exposure to stressors has hardly received any attention. In this work, Western honeybees (Apis mellifera) were sequentially or simultaneously infected by the microsporidian parasite Nosema ceranae and chronically exposed to a sublethal dose of the insecticide fipronil, respectively chosen as biological and chemical stressors. Interestingly, every combination tested led to a synergistic effect on honeybee survival, with the most significant impacts when stressors were applied at the emergence of honeybees. Our study presents significant outcomes on beekeeping management but also points out the potential risks incurred by any living organism frequently exposed to both pathogens and insecticides in their habitat

    Systematic discovery of unannotated genes in 11 yeast species using a database of orthologous genomic segments

    Get PDF
    <p>Abstract</p> <p>Background</p> <p>In standard BLAST searches, no information other than the sequences of the query and the database entries is considered. However, in situations where two genes from different species have only borderline similarity in a BLAST search, the discovery that the genes are located within a region of conserved gene order (synteny) can provide additional evidence that they are orthologs. Thus, for interpreting borderline search results, it would be useful to know whether the syntenic context of a database hit is similar to that of the query. This principle has often been used in investigations of particular genes or genomic regions, but to our knowledge it has never been implemented systematically.</p> <p>Results</p> <p>We made use of the synteny information contained in the Yeast Gene Order Browser database for 11 yeast species to carry out a systematic search for protein-coding genes that were overlooked in the original annotations of one or more yeast genomes but which are syntenic with their orthologs. Such genes tend to have been overlooked because they are short, highly divergent, or contain introns. The key features of our software - called SearchDOGS - are that the database entries are classified into sets of genomic segments that are already known to be orthologous, and that very weak BLAST hits are retained for further analysis if their genomic location is similar to that of the query. Using SearchDOGS we identified 595 additional protein-coding genes among the 11 yeast species, including two new genes in <it>Saccharomyces cerevisiae</it>. We found additional genes for the mating pheromone a-factor in six species including <it>Kluyveromyces lactis</it>.</p> <p>Conclusions</p> <p>SearchDOGS has proven highly successful for identifying overlooked genes in the yeast genomes. We anticipate that our approach can be adapted for study of further groups of species, such as bacterial genomes. More generally, the concept of doing sequence similarity searches against databases to which external information has been added may prove useful in other settings.</p

    A novel mitochondrial DnaJ/Hsp40 family protein BIL2 promotes plant growth and resistance against environmental stress in brassinosteroid signaling

    Get PDF
    Funding Information: Acknowledgments We thank Dr. Tsuyoshi Nakagawa (Shimane University) for the gift of the gateway vectors, pGWB2, pGWB80, pGWB5, and pGWB3. This work was supported in part by funding from the Program for Promotion of Basic Research Activities for Innovation Bioscience (PROBRAIN) to T.N. and T.A., and CREST, Japan Science and Technology Agency to T.N. and T.A.Plant steroid hormones, brassinosteroids, are essential for growth, development and responses to environmental stresses in plants. Although BR signaling proteins are localized in many organelles, i.e., the plasma membrane, nuclei, endoplasmic reticulum and vacuole, the details regarding the BR signaling pathway from perception at the cellular membrane receptor BRASSINOSTEROID INSENSITIVE 1 (BRI1) to nuclear events include several steps. Brz (Brz220) is a specific inhibitor of BR biosynthesis. In this study, we used Brz-mediated chemical genetics to identify Brz-insensitive-long hypocotyls 2-1D (bil2-1D). The BIL2 gene encodes a mitochondrial-localized DnaJ/Heat shock protein 40 (DnaJ/Hsp40) family, which is involved in protein folding. BIL2-overexpression plants (BIL2-OX) showed cell elongation under Brz treatment, increasing the growth of plant inflorescence and roots, the regulation of BR-responsive gene expression and suppression against the dwarfed BRI1-deficient mutant. BIL2-OX also showed resistance against the mitochondrial ATPase inhibitor oligomycin and higher levels of exogenous ATP compared with wild-type plants. BIL2 participates in resistance against salinity stress and strong light stress. Our results indicate that BIL2 induces cell elongation during BR signaling through the promotion of ATP synthesis in mitochondria.Peer reviewe

    A Complete Sequence and Transcriptomic Analyses of Date Palm (Phoenix dactylifera L.) Mitochondrial Genome

    Get PDF
    Based on next-generation sequencing data, we assembled the mitochondrial (mt) genome of date palm (Phoenix dactylifera L.) into a circular molecule of 715,001 bp in length. The mt genome of P. dactylifera encodes 38 proteins, 30 tRNAs, and 3 ribosomal RNAs, which constitute a gene content of 6.5% (46,770 bp) over the full length. The rest, 93.5% of the genome sequence, is comprised of cp (chloroplast)-derived (10.3% with respect to the whole genome length) and non-coding sequences. In the non-coding regions, there are 0.33% tandem and 2.3% long repeats. Our transcriptomic data from eight tissues (root, seed, bud, fruit, green leaf, yellow leaf, female flower, and male flower) showed higher gene expression levels in male flower, root, bud, and female flower, as compared to four other tissues. We identified 120 potential SNPs among three date palm cultivars (Khalas, Fahal, and Sukry), and successfully found seven SNPs in the coding sequences. A phylogenetic analysis, based on 22 conserved genes of 15 representative plant mitochondria, showed that P. dactylifera positions at the root of all sequenced monocot mt genomes. In addition, consistent with previous discoveries, there are three co-transcribed gene clusters–18S-5S rRNA, rps3-rpl16 and nad3-rps12–in P. dactylifera, which are highly conserved among all known mitochondrial genomes of angiosperms

    Molecular Phylogeny and Evolution of Parabasalia with Improved Taxon Sampling and New Protein Markers of Actin and Elongation Factor-1α

    Get PDF
    BACKGROUND: Inferring the evolutionary history of phylogenetically isolated, deep-branching groups of taxa-in particular determining the root-is often extraordinarily difficult because their close relatives are unavailable as suitable outgroups. One of these taxonomic groups is the phylum Parabasalia, which comprises morphologically diverse species of flagellated protists of ecological, medical, and evolutionary significance. Indeed, previous molecular phylogenetic analyses of members of this phylum have yielded conflicting and possibly erroneous inferences. Furthermore, many species of Parabasalia are symbionts in the gut of termites and cockroaches or parasites and therefore formidably difficult to cultivate, rendering available data insufficient. Increasing the numbers of examined taxa and informative characters (e.g., genes) is likely to produce more reliable inferences. PRINCIPAL FINDINGS: Actin and elongation factor-1α genes were identified newly from 22 species of termite-gut symbionts through careful manipulations and seven cultured species, which covered major lineages of Parabasalia. Their protein sequences were concatenated and analyzed with sequences of previously and newly identified glyceraldehyde-3-phosphate dehydrogenase and the small-subunit rRNA gene. This concatenated dataset provided more robust phylogenetic relationships among major groups of Parabasalia and a more plausible new root position than those previously reported. CONCLUSIONS/SIGNIFICANCE: We conclude that increasing the number of sampled taxa as well as the addition of new sequences greatly improves the accuracy and robustness of the phylogenetic inference. A morphologically simple cell is likely the ancient form in Parabasalia as opposed to a cell with elaborate flagellar and cytoskeletal structures, which was defined as most basal in previous inferences. Nevertheless, the evolution of Parabasalia is complex owing to several independent multiplication and simplification events in these structures. Therefore, systematics based solely on morphology does not reflect the evolutionary history of parabasalids

    Different rates of spontaneous mutation of chloroplastic and nuclear viroids as determined by high-fidelity ultra-deep sequencing

    Full text link
    [EN] Mutation rates vary by orders of magnitude across biological systems, being higher for simpler genomes. The simplest known genomes correspond to viroids, subviral plant replicons constituted by circular non-coding RNAs of few hundred bases. Previous work has revealed an extremely high mutation rate for chrysanthemum chlorotic mottle viroid, a chloroplastreplicating viroid. However, whether this is a general feature of viroids remains unclear. Here, we have used high-fidelity ultra-deep sequencing to determine the mutation rate in a common host (eggplant) of two viroids, each representative of one family: the chloroplastic eggplant latent viroid (ELVd, Avsunviroidae) and the nuclear potato spindle tuber viroid (PSTVd, Pospiviroidae). This revealed higher mutation frequencies in ELVd than in PSTVd, as well as marked differences in the types of mutations produced. Rates of spontaneous mutation, quantified in vivo using the lethal mutation method, ranged from 1/1000 to 1/800 for ELVd and from 1/7000 to 1/3800 for PSTVd depending on sequencing run. These results suggest that extremely high mutability is a common feature of chloroplastic viroids, whereas the mutation rates of PSTVd and potentially other nuclear viroids appear significantly lower and closer to those of some RNA viruses.This work was supported by the European Research Council (erc.europa.eu; ERC-2011-StG-281191-VIRMUT to RS), the Spanish Ministerio de Economia y Competitividad (www.mineco.gob.es; BFU2013-41329 grant to RS, BFU2014-56812-P grant to RF, and a predoctoral fellowship to ALC), and the Spanish Junta de Comunidades de Castilla-La Mancha (www.castillalamancha.es;postdoctoral fellowship to CB). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.López-Carrasco, MA.; Ballesteros Martínez, C.; Sentandreu, V.; Delgado Villar, SG.; Gago Zachert, SP.; Flores Pedauye, R.; Sanjuan Verdeguer, R. (2017). Different rates of spontaneous mutation of chloroplastic and nuclear viroids as determined by high-fidelity ultra-deep sequencing. PLoS Pathogens. 13(9):1-17. https://doi.org/10.1371/journal.ppat.1006547S117139Ganai, R. A., & Johansson, E. (2016). DNA Replication—A Matter of Fidelity. Molecular Cell, 62(5), 745-755. doi:10.1016/j.molcel.2016.05.003Lynch, M. (2010). Evolution of the mutation rate. Trends in Genetics, 26(8), 345-352. doi:10.1016/j.tig.2010.05.003Sanjuán, R., & Domingo-Calap, P. (2016). Mechanisms of viral mutation. Cellular and Molecular Life Sciences, 73(23), 4433-4448. doi:10.1007/s00018-016-2299-6Gago, S., Elena, S. F., Flores, R., & Sanjuan, R. (2009). Extremely High Mutation Rate of a Hammerhead Viroid. Science, 323(5919), 1308-1308. doi:10.1126/science.1169202Flores, R., Gago-Zachert, S., Serra, P., Sanjuán, R., & Elena, S. F. (2014). Viroids: Survivors from the RNA World? Annual Review of Microbiology, 68(1), 395-414. doi:10.1146/annurev-micro-091313-103416Flores, R., Minoia, S., Carbonell, A., Gisel, A., Delgado, S., López-Carrasco, A., … Di Serio, F. (2015). Viroids, the simplest RNA replicons: How they manipulate their hosts for being propagated and how their hosts react for containing the infection. Virus Research, 209, 136-145. doi:10.1016/j.virusres.2015.02.027Steger, G., & Perreault, J.-P. (2016). Structure and Associated Biological Functions of Viroids. Advances in Virus Research, 141-172. doi:10.1016/bs.aivir.2015.11.002Diener, T. O. (1989). Circular RNAs: relics of precellular evolution? Proceedings of the National Academy of Sciences, 86(23), 9370-9374. doi:10.1073/pnas.86.23.9370Ambrós, S., Hernández, C., & Flores, R. (1999). Rapid generation of genetic heterogeneity in progenies from individual cDNA clones of peach latent mosaic viroid in its natural host The data reported in this paper are in the EMBL nucleotide sequence database and assigned the accession nos AJ241818–AJ241850. Journal of General Virology, 80(8), 2239-2252. doi:10.1099/0022-1317-80-8-2239Navarro, J.-A., Vera, A., & Flores, R. (2000). A Chloroplastic RNA Polymerase Resistant to Tagetitoxin Is Involved in Replication of Avocado Sunblotch Viroid. Virology, 268(1), 218-225. doi:10.1006/viro.1999.0161Rodio, M.-E., Delgado, S., De Stradis, A., Gómez, M.-D., Flores, R., & Di Serio, F. (2007). A Viroid RNA with a Specific Structural Motif Inhibits Chloroplast Development. The Plant Cell, 19(11), 3610-3626. doi:10.1105/tpc.106.049775Carbonell, A., De la Peña, M., Flores, R., & Gago, S. (2006). Effects of the trinucleotide preceding the self-cleavage site on eggplant latent viroid hammerheads: differences in co- and post-transcriptional self-cleavage may explain the lack of trinucleotide AUC in most natural hammerheads. Nucleic Acids Research, 34(19), 5613-5622. doi:10.1093/nar/gkl717Hutchins, C. J., Rathjen, P. D., Forster, A. C., & Symons, R. H. (1986). Self-cleavage of plus and minus RNA transcripts of avocado sunblotch viroid. Nucleic Acids Research, 14(9), 3627-3640. doi:10.1093/nar/14.9.3627PRODY, G. A., BAKOS, J. T., BUZAYAN, J. M., SCHNEIDER, I. R., & BRUENING, G. (1986). Autolytic Processing of Dimeric Plant Virus Satellite RNA. Science, 231(4745), 1577-1580. doi:10.1126/science.231.4745.1577Nohales, M.-A., Molina-Serrano, D., Flores, R., & Daros, J.-A. (2012). Involvement of the Chloroplastic Isoform of tRNA Ligase in the Replication of Viroids Belonging to the Family Avsunviroidae. Journal of Virology, 86(15), 8269-8276. doi:10.1128/jvi.00629-12Branch, A. D., Benenfeld, B. J., & Robertson, H. D. (1988). Evidence for a single rolling circle in the replication of potato spindle tuber viroid. Proceedings of the National Academy of Sciences, 85(23), 9128-9132. doi:10.1073/pnas.85.23.9128Daros, J.-A., & Flores, R. (2004). Arabidopsis thaliana has the enzymatic machinery for replicating representative viroid species of the family Pospiviroidae. Proceedings of the National Academy of Sciences, 101(17), 6792-6797. doi:10.1073/pnas.0401090101Feldstein, P. A., Hu, Y., & Owens, R. A. (1998). Precisely full length, circularizable, complementary RNA: An infectious form of potato spindle tuber viroid. Proceedings of the National Academy of Sciences, 95(11), 6560-6565. doi:10.1073/pnas.95.11.6560Gas, M.-E., Hernández, C., Flores, R., & Daròs, J.-A. (2007). Processing of Nuclear Viroids In Vivo: An Interplay between RNA Conformations. PLoS Pathogens, 3(11), e182. doi:10.1371/journal.ppat.0030182Nohales, M.-A., Flores, R., & Daros, J.-A. (2012). Viroid RNA redirects host DNA ligase 1 to act as an RNA ligase. Proceedings of the National Academy of Sciences, 109(34), 13805-13810. doi:10.1073/pnas.1206187109Brass, J. R. J., Owens, R. A., Matoušek, J., & Steger, G. (2017). Viroid quasispecies revealed by deep sequencing. RNA Biology, 14(3), 317-325. doi:10.1080/15476286.2016.1272745Bull, J. J., Sanjuán, R., & Wilke, C. O. (2007). Theory of Lethal Mutagenesis for Viruses. Journal of Virology, 81(6), 2930-2939. doi:10.1128/jvi.01624-06Cuevas, J. M., González-Candelas, F., Moya, A., & Sanjuán, R. (2009). Effect of Ribavirin on the Mutation Rate and Spectrum of Hepatitis C Virus In Vivo. Journal of Virology, 83(11), 5760-5764. doi:10.1128/jvi.00201-09Ribeiro, R. M., Li, H., Wang, S., Stoddard, M. B., Learn, G. H., Korber, B. T., … Perelson, A. S. (2012). Quantifying the Diversification of Hepatitis C Virus (HCV) during Primary Infection: Estimates of the In Vivo Mutation Rate. PLoS Pathogens, 8(8), e1002881. doi:10.1371/journal.ppat.1002881Acevedo, A., Brodsky, L., & Andino, R. (2013). Mutational and fitness landscapes of an RNA virus revealed through population sequencing. Nature, 505(7485), 686-690. doi:10.1038/nature12861Cuevas, J. M., Geller, R., Garijo, R., López-Aldeguer, J., & Sanjuán, R. (2015). Extremely High Mutation Rate of HIV-1 In Vivo. PLOS Biology, 13(9), e1002251. doi:10.1371/journal.pbio.1002251Acevedo, A., & Andino, R. (2014). Library preparation for highly accurate population sequencing of RNA viruses. Nature Protocols, 9(7), 1760-1769. doi:10.1038/nprot.2014.118Kennedy, S. R., Schmitt, M. W., Fox, E. J., Kohrn, B. F., Salk, J. J., Ahn, E. H., … Loeb, L. A. (2014). Detecting ultralow-frequency mutations by Duplex Sequencing. Nature Protocols, 9(11), 2586-2606. doi:10.1038/nprot.2014.170Franklin, R. M. (1966). Purification and properties of the replicative intermediate of the RNA bacteriophage R17. Proceedings of the National Academy of Sciences, 55(6), 1504-1511. doi:10.1073/pnas.55.6.1504López-Carrasco, A., Gago-Zachert, S., Mileti, G., Minoia, S., Flores, R., & Delgado, S. (2015). The transcription initiation sites of eggplant latent viroid strands map within distinct motifs in theirin vivoRNA conformations. RNA Biology, 13(1), 83-97. doi:10.1080/15476286.2015.1119365Keese, P., & Symons, R. H. (1985). Domains in viroids: evidence of intermolecular RNA rearrangements and their contribution to viroid evolution. Proceedings of the National Academy of Sciences, 82(14), 4582-4586. doi:10.1073/pnas.82.14.4582López-Carrasco, A., & Flores, R. (2016). Dissecting the secondary structure of the circular RNA of a nuclear viroid in vivo: A «naked» rod-like conformation similar but not identical to that observed in vitro. RNA Biology, 14(8), 1046-1054. doi:10.1080/15476286.2016.1223005Flores, R., Hernandez, C., de la Peña, M., Vera, A., & Daros, J.-A. (2001). Hammerhead Ribozyme Structure and Function in Plant RNA Replication. Ribonucleases - Part A, 540-552. doi:10.1016/s0076-6879(01)41175-xMartick, M., & Scott, W. G. (2006). Tertiary Contacts Distant from the Active Site Prime a Ribozyme for Catalysis. Cell, 126(2), 309-320. doi:10.1016/j.cell.2006.06.036Ruffner, D. E., Stormo, G. D., & Uhlenbeck, O. C. (1990). Sequence requirements of the hammerhead RNA self-cleavage reaction. Biochemistry, 29(47), 10695-10702. doi:10.1021/bi00499a018Flores, R., Serra, P., Minoia, S., Di Serio, F., & Navarro, B. (2012). Viroids: From Genotype to Phenotype Just Relying on RNA Sequence and Structural Motifs. Frontiers in Microbiology, 3. doi:10.3389/fmicb.2012.00217Owens, R. A., Chen, W., Hu, Y., & Hsu, Y.-H. (1995). Suppression of Potato Spindle Tuber Viroid Replication and Symptom Expression by Mutations Which Stabilize the Pathogenicity Domain. Virology, 208(2), 554-564. doi:10.1006/viro.1995.1186Takeda, R., Petrov, A. I., Leontis, N. B., & Ding, B. (2011). A Three-Dimensional RNA Motif in Potato spindle tuber viroid Mediates Trafficking from Palisade Mesophyll to Spongy Mesophyll in Nicotiana benthamiana. The Plant Cell, 23(1), 258-272. doi:10.1105/tpc.110.081414Zhong, X., Leontis, N., Qian, S., Itaya, A., Qi, Y., Boris-Lawrie, K., & Ding, B. (2006). Tertiary Structural and Functional Analyses of a Viroid RNA Motif by Isostericity Matrix and Mutagenesis Reveal Its Essential Role in Replication. Journal of Virology, 80(17), 8566-8581. doi:10.1128/jvi.00837-06Zhong, X., Tao, X., Stombaugh, J., Leontis, N., & Ding, B. (2007). Tertiary structure and function of an RNA motif required for plant vascular entry to initiate systemic trafficking. The EMBO Journal, 26(16), 3836-3846. doi:10.1038/sj.emboj.7601812Zhong, X., Archual, A. J., Amin, A. A., & Ding, B. (2008). A Genomic Map of Viroid RNA Motifs Critical for Replication and Systemic Trafficking. The Plant Cell, 20(1), 35-47. doi:10.1105/tpc.107.056606Thomas, M. J., Platas, A. A., & Hawley, D. K. (1998). Transcriptional Fidelity and Proofreading by RNA Polymerase II. Cell, 93(4), 627-637. doi:10.1016/s0092-8674(00)81191-5Gout, J.-F., Thomas, W. K., Smith, Z., Okamoto, K., & Lynch, M. (2013). Large-scale detection of in vivo transcription errors. Proceedings of the National Academy of Sciences, 110(46), 18584-18589. doi:10.1073/pnas.1309843110Hedtke, B. (1997). Mitochondrial and Chloroplast Phage-Type RNA Polymerases in Arabidopsis. Science, 277(5327), 809-811. doi:10.1126/science.277.5327.809Lerbs-Mache, S. (1993). The 110-kDa polypeptide of spinach plastid DNA-dependent RNA polymerase: single-subunit enzyme or catalytic core of multimeric enzyme complexes? Proceedings of the National Academy of Sciences, 90(12), 5509-5513. doi:10.1073/pnas.90.12.5509Oldenkott, B., Yamaguchi, K., Tsuji-Tsukinoki, S., Knie, N., & Knoop, V. (2014). Chloroplast RNA editing going extreme: more than 3400 events of C-to-U editing in the chloroplast transcriptome of the lycophyteSelaginella uncinata. RNA, 20(10), 1499-1506. doi:10.1261/rna.045575.114Codoñer, F. M., Darós, J.-A., Solé, R. V., & Elena, S. F. (2006). The Fittest versus the Flattest: Experimental Confirmation of the Quasispecies Effect with Subviral Pathogens. PLoS Pathogens, 2(12), e136. doi:10.1371/journal.ppat.0020136Eigen, M. (1971). Selforganization of matter and the evolution of biological macromolecules. Die Naturwissenschaften, 58(10), 465-523. doi:10.1007/bf00623322Lynch, M. (2011). The Lower Bound to the Evolution of Mutation Rates. Genome Biology and Evolution, 3, 1107-1118. doi:10.1093/gbe/evr066Bradwell, K., Combe, M., Domingo-Calap, P., & Sanjuán, R. (2013). Correlation Between Mutation Rate and Genome Size in Riboviruses: Mutation Rate of Bacteriophage Qβ. Genetics, 195(1), 243-251. doi:10.1534/genetics.113.154963Drake, J. W. (1991). A constant rate of spontaneous mutation in DNA-based microbes. Proceedings of the National Academy of Sciences, 88(16), 7160-7164. doi:10.1073/pnas.88.16.7160Schmitt, M. W., Kennedy, S. R., Salk, J. J., Fox, E. J., Hiatt, J. B., & Loeb, L. A. (2012). Detection of ultra-rare mutations by next-generation sequencing. Proceedings of the National Academy of Sciences, 109(36), 14508-14513. doi:10.1073/pnas.120871510
    corecore