194 research outputs found

    How the other half lives: CRISPR-Cas's influence on bacteriophages

    Full text link
    CRISPR-Cas is a genetic adaptive immune system unique to prokaryotic cells used to combat phage and plasmid threats. The host cell adapts by incorporating DNA sequences from invading phages or plasmids into its CRISPR locus as spacers. These spacers are expressed as mobile surveillance RNAs that direct CRISPR-associated (Cas) proteins to protect against subsequent attack by the same phages or plasmids. The threat from mobile genetic elements inevitably shapes the CRISPR loci of archaea and bacteria, and simultaneously the CRISPR-Cas immune system drives evolution of these invaders. Here we highlight our recent work, as well as that of others, that seeks to understand phage mechanisms of CRISPR-Cas evasion and conditions for population coexistence of phages with CRISPR-protected prokaryotes.Comment: 24 pages, 8 figure

    Comparative genomics of the class 4 histone deacetylase family indicates a complex evolutionary history

    Get PDF
    BACKGROUND: Histone deacetylases are enzymes that modify core histones and play key roles in transcriptional regulation, chromatin assembly, DNA repair, and recombination in eukaryotes. Three types of related histone deacetylases (classes 1, 2, and 4) are widely found in eukaryotes, and structurally related proteins have also been found in some prokaryotes. Here we focus on the evolutionary history of the class 4 histone deacetylase family. RESULTS: Through sequence similarity searches against sequenced genomes and expressed sequence tag data, we identified members of the class 4 histone deacetylase family in 45 eukaryotic and 37 eubacterial species representative of very distant evolutionary lineages. Multiple phylogenetic analyses indicate that the phylogeny of these proteins is, in many respects, at odds with the phylogeny of the species in which they are found. In addition, the eukaryotic members of the class 4 histone deacetylase family clearly display an anomalous phyletic distribution. CONCLUSION: The unexpected phylogenetic relationships within the class 4 histone deacetylase family and the anomalous phyletic distribution of these proteins within eukaryotes might be explained by two mechanisms: ancient gene duplication followed by differential gene losses and/or horizontal gene transfer. We discuss both possibilities in this report, and suggest that the evolutionary history of the class 4 histone deacetylase family may have been shaped by horizontal gene transfers

    Reduced stability of mRNA secondary structure near the translation-initiation site in dsDNA viruses

    Get PDF
    <p>Abstract</p> <p>Background</p> <p>Recent studies have demonstrated a selection pressure for reduced mRNA secondary-structure stability near the start codon of coding sequences. This selection pressure can be observed in bacteria, archaea, and eukaryotes, and is likely caused by the requirement of efficient translation initiation in cellular organism.</p> <p>Results</p> <p>Here, we surveyed the complete genomes of 650 dsDNA virus strains for signals of reduced stability of mRNA secondary structure near the start codon. Our analysis included viruses infecting eukaryotic, prokaryotic, and archaeic hosts. We found that many viruses showed evidence for reduced mRNA secondary-structure stability near the start codon. The effect was most pronounced in viruses infecting prokaryotes, but was also observed in viruses infecting eukaryotes and archaea. The reduction in stability generally increased with increasing genomic GC content. For bacteriophage, the reduction was correlated with a corresponding reduction of stability in the phage hosts.</p> <p>Conclusions</p> <p>We conclude that reduced stability of the mRNA secondary structure near the start codon is a common feature for dsDNA viruses, likely driven by the same selective pressures that cause it in cellular organisms.</p

    Domain-Domain Interactions Underlying Herpesvirus-Human Protein-Protein Interaction Networks

    Get PDF
    Protein-domains play an important role in mediating protein-protein interactions. Furthermore, the same domain-pairs mediate different interactions in different contexts and in various organisms, and therefore domain-pairs are considered as the building blocks of interactome networks. Here we extend these principles to the host-virus interface and find the domain-pairs that potentially mediate human-herpesvirus interactions. Notably, we find that the same domain-pairs used by other organisms for mediating their interactions underlie statistically significant fractions of human-virus protein inter-interaction networks. Our analysis shows that viral domains tend to interact with human domains that are hubs in the human domain-domain interaction network. This may enable the virus to easily interfere with a variety of mechanisms and processes involving various and different human proteins carrying the relevant hub domain. Comparative genomics analysis provides hints at a molecular mechanism by which the virus acquired some of its interacting domains from its human host

    BRCA1 Interacts with Smad3 and Regulates Smad3-Mediated TGF-β Signaling during Oxidative Stress Responses

    Get PDF
    BRCA1 is a key regulatory protein participating in cell cycle checkpoint and DNA damage repair networks. BRCA1 plays important roles in protecting numerous cellular processes in response to cell damaging signals. Transforming growth factor-beta (TGF-beta) is a potent regulator of growth, apoptosis and invasiveness of tumor cells. TFG-beta activates Smad signaling via its two cell surface receptors, the TbetaRII and ALK5/TbetaRI, leading to Smad-mediated transcriptional regulation.Here, we report an important role of BRCA1 in modulating TGF-beta signaling during oxidative stress responses. Wild-type (WT) BRCA1, but not mutated BRCA1 failed to activate TGF-beta mediated transactivation of the TGF-beta responsive reporter, p3TP-Lux. Further, WT-BRCA1, but not mutated BRCA1 increased the expression of Smad3 protein in a dose-dependent manner, while silencing of WT-BRCA1 by siRNA decreased Smad3 and Smad4 interaction induced by TGF-beta in MCF-7 breast cancer cells. BRCA1 interacted with Smad3 upon TGF-beta1 stimulation in MCF-7 cells and this interaction was mediated via the domain of 298-436aa of BRCA1 and Smad3 domain of 207-426aa. In addition, H(2)O(2) increased the colocalization and the interaction of Smad3 with WT-BRCA1. Interestingly, TGF-beta1 induced Smad3 and Smad4 interaction was increased in the presence of H(2)O(2) in cells expressing WT-BRCA1, while the TGF-beta1 induced interaction between Smad3 and Smad4 was decreased upon H(2)O(2) treatment in a dose-dependent manner in HCC1937 breast cancer cells, deficient for endogenous BRCA1. This interaction between Smad3 and Smad4 was increased in reconstituted HCC1937 cells expressing WT-BRCA1 (HCC1937/BRCA1). Further, loss of BRCA1 resulted in H(2)O(2) induced nuclear export of phosphor-Smad3 protein to the cytoplasm, resulting decreased of Smad3 and Smad4 interaction induced by TGF-beta and in significant decrease in Smad3 and Smad4 transcriptional activities.These results strongly suggest that loss or reduction of BRCA1 alters TGF-beta growth inhibiting activity via Smad3 during oxidative stress responses

    Structural Organization of DNA in Chlorella Viruses

    Get PDF
    Chlorella viruses have icosahedral capsids with an internal membrane enclosing their large dsDNA genomes and associated proteins. Their genomes are packaged in the particles with a predicted DNA density of ca. 0.2 bp nm−3. Occasionally infection of an algal cell by an individual particle fails and the viral DNA is dynamically ejected from the capsid. This shows that the release of the DNA generates a force, which can aid in the transfer of the genome into the host in a successful infection. Imaging of ejected viral DNA indicates that it is intimately associated with proteins in a periodic fashion. The bulk of the protein particles detected by atomic force microscopy have a size of ∼60 kDa and two proteins (A278L and A282L) of about this size are among 6 basic putative DNA binding proteins found in a proteomic analysis of DNA binding proteins packaged in the virion. A combination of fluorescence images of ejected DNA and a bioinformatics analysis of the DNA reveal periodic patterns in the viral DNA. The periodic distribution of GC rich regions in the genome provides potential binding sites for basic proteins. This DNA/protein aggregation could be responsible for the periodic concentration of fluorescently labeled DNA observed in ejected viral DNA. Collectively the data indicate that the large chlorella viruses have a DNA packaging strategy that differs from bacteriophages; it involves proteins and share similarities to that of chromatin structure in eukaryotes

    The Universal Plausibility Metric (UPM) & Principle (UPP)

    Get PDF
    <p>Abstract</p> <p>Background</p> <p>Mere possibility is not an adequate basis for asserting scientific plausibility. A precisely defined universal bound is needed beyond which the assertion of <it>plausibility</it>, particularly in life-origin models, can be considered operationally falsified. But can something so seemingly relative and subjective as plausibility ever be quantified? Amazingly, the answer is, "Yes." A method of objectively measuring the plausibility of any chance hypothesis (The Universal Plausibility Metric [UPM]) is presented. A numerical inequality is also provided whereby any chance hypothesis can be definitively falsified when its UPM metric of ξ is < 1 (The Universal Plausibility Principle [UPP]). Both UPM and UPP pre-exist and are independent of any experimental design and data set.</p> <p>Conclusion</p> <p>No low-probability hypothetical plausibility assertion should survive peer-review without subjection to the UPP inequality standard of formal falsification (ξ < 1).</p

    Changes in the composition of the RNA virome mark evolutionary transitions in green plants

    Get PDF
    Background: The known plant viruses mostly infect angiosperm hosts and have RNA or small DNA genomes. The only other lineage of green plants with a relatively well-studied virome, unicellular chlorophyte algae, is mostly infected by viruses with large DNA genomes. Thus RNA viruses and small DNA viruses seem to completely displace large DNA virus genomes in late branching angiosperms. To understand better the expansion of RNA viruses in the taxonomic span between algae and angiosperms, we analyzed the transcriptomes of 66 non-angiosperm plants characterized by the 1000 Plants Genomes Project. Results: We found homologs of virus RNA-dependent RNA polymerases in 28 non-angiosperm plant species, including algae, mosses, liverworts (Marchantiophyta), hornworts (Anthocerotophyta), lycophytes, a horsetail Equisetum, and gymnosperms. Polymerase genes in algae were most closely related to homologs from double-stranded RNA viruses leading latent or persistent lifestyles. Land plants, in addition, contained polymerases close to the homologs from single-stranded RNA viruses of angiosperms, capable of productive infection and systemic spread. For several polymerases, a cognate capsid protein was found in the same library. Another virus hallmark gene family, encoding the 30 K movement proteins, was found in lycophytes and monilophytes but not in mosses or algae. Conclusions: The broadened repertoire of RNA viruses suggests that colonization of land and growth in anatomical complexity in land plants coincided with the acquisition of novel sets of viruses with different strategies of infection and reproduction.We thank the colleagues at the 1000 Plant Genomes Project for helping us to access the transcriptomes used in this study via the iPlant Collaborative. We are grateful to Javier Forment (IBMCP-CSIC), Vincent Lefort (PhyML), and the E-Biothon team (E-Biothon platform is supported by CNRS, IBM, INRIA, l'Institut Francais de Bioinformatique and SysFera) for expert help with high-performance computing; to Yuri Wolf, Jan Kreuze, Eddie Holmes, and Mang Shi for sharing sequence data and alignments; to Sejo Sabanadzovic, Jan Kreuze, and the anonymous reviewers for helpful virtual discussions and critical remarks; and to Natalia Mushegian for technical assistance. SFF was supported by grants BFU2015-65037P from Spain Ministry of Economy and Competitiveness and PROMETEOII/2014/021 from Generalitat Valenciana. ARM is a Program Director at the US National Science Foundation (NSF); his work on this project was supported by the NSF Independent Research and Development Program, but the statements and opinions expressed herein are made in the personal capacity and do not constitute the endorsement by NSF or the government of the United States.Mushegian, A.; Shipunov, A.; Elena Fito, SF. (2016). Changes in the composition of the RNA virome mark evolutionary transitions in green plants. BMC Biology. 14(68):1-14. https://doi.org/10.1186/s12915-016-0288-8S1141468Roossinck MJ. Plant virus metagenomics: biodiversity and ecology. Annu Rev Genet. 2012;46:357–67.Koonin EV, Dolja VV, Krupovic M. Origins and evolution of viruses of eukaryotes: The ultimate modularity. Virology. 2015;479-480:2–25.Yamada T, Onimatsu H, Van Etten JL. Chlorella viruses. Adv Virus Res. 2006;66:293–336.Van Etten JL, Dunigan DD. Chloroviruses: not your everyday plant virus. Trends Plant Sci. 2012;17:1–8.Zhang T, Jiang Y, Dong W. A novel monopartite dsRNA virus isolated from the phytopathogenic fungus Ustilaginoidea virens and ancestrally related to a mitochondria-associated dsRNA in the green alga Bryopsis. Virology. 2014;462:227–35.Wilson WH, Van Etten JL, Allen MJ. The Phycodnaviridae: the story of how tiny giants rule the world. Curr Top Microbiol Immunol. 2009;328:1–42.Iyer LM, Balaji S, Koonin EV, Aravind L. Evolutionary genomics of nucleo-cytoplasmic large DNA viruses. Virus Res. 2006;117:156–84.Colson P, De Lamballerie X, Yutin N, Asgari S, Bigot Y, Bideshi DK, et al. “Megavirales”, a proposed new order for eukaryotic nucleocytoplasmic large DNA viruses. Arch Virol. 2013;158:2517–21.Gibbs AJ, Torronen M, Mackenzie AM, Wood 2nd JT, Amstrong JS, Kondo H, et al. The enigmatic genome of Chara australis virus. J Gen Virol. 2011;92:2679–90.Valverde RA, Sabanadzovic S. A novel plant virus with unique properties infecting Japanese holly fern. J Gen Virol. 2009;90:42–9.Han SS, Karasev AV, Ieki H, Iwanami T. Nucleotide sequence and taxonomy of Cycas necrotic stunt virus. Brief report. Arch Virol. 2002;147:2207–14.Lockhart B, Fetzer JL, Olszewski NE. Preliminary characterization of Cycad leaf necrosis virus, the first badnavirus identified in cycads. Phytopathology. 2006;96:S70.Mushegian AR, Elena SF. Evolution of plant virus movement proteins from the 30 K superfamily and of their homologs integrated in plant genomes. Virology. 2015;476:304–15.Maumus F, Epert A, Nogué F, Blanc G. Plant genomes enclose footprints of past infections by giant virus relatives. Nat Commun. 2014;5:4268.Wickett NJ, Mirarab S, Nguyen N, Warnow T, Carpenter E, Matasci N, et al. Phylotranscriptomic analysis of the origin and early diversification of land plants. Proc Natl Acad Sci USA. 2014;111:E4859–68.Merchant N, Lyons E, Goff S, Vaughn M, Ware D, Micklos D, et al. The iPlant collaborative: cyberinfrastructure for enabling data to discovery for the life sciences. PLoS Biol. 2016;14:e1002342.Becker B, Marin B. Streptophyte algae and the origin of embryophytes. Ann Bot. 2009;103:999–1004.Koonin EV, Dolja VV. A virocentric perspective on the evolution of life. Curr Op Virol. 2013;3:546–57.Matasci N, Huang LH, Yan Z, Carpenter EJ, Wickett NJ, Mirarab S, et al. Data access for the 1000 plants (1KP) project. Gigascience. 2014;3:17.Makarova KS, Aravind L, Grishin NV, Rogozin IB, Koonin EV. A DNA repair system specific for thermophilic Archaea and bacteria predicted by genomic context analysis. Nucleic Acids Res. 2002;30:482–96.Majumdar I, Kinch LN, Grishin NV. A database of domain definitions for proteins with complex interdomain geometry. PLoS ONE. 2009;4:e5084.Černý J, Černá Bolfíková B, Zanotto PMA, Grubhoffer L, Růžek D. A deep phylogeny of viral and cellular right-hand polymerases. Infec Genet Evol. 2015;36:275–86.Iyer LM, Koonin EV, Aravind L. Evolutionary connection between the catalytic subunits of DNA-dependent RNA polymerases and eukaryotic RNA-dependent RNA polymerases and the origin of RNA polymerases. BMC Struct Biol. 2003;3:1.Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25:3389–402.Koga R, Horiuchi H, Fukuhara T. Double-stranded RNA replicons associated with chloroplasts of a green alga, Bryopsis cinicola. Plant Mol Biol. 2003;51:991–9.Jablonski SA, Morrow CD. Mutation of the aspartic acid residues of the GDD sequence motif of poliovirus RNA-dependent RNA polymerase results in enzymes with altered metal ion requirements for activity. J Virol. 1995;69:1532–9.Morin B, Whelan SPJ. La protéine L des Mononegavirales. Virologie. 2012;16:258–68.Liang B, Li Z, Jenni S, Rahmeh AA, Morin BM, Grant T, et al. Structure of the L protein of Vesicular stomatitis virus from electron cryomicroscopy. Cell. 2015;162:314–27.Chabannes M, Iskra-Caruana ML. Endogenous pararetroviruses – a reservoir of virus infection in plants. Curr Opin Virol. 2013;3:615–20.Bousios A, Darzentas N. Sirevirus LTR retrotransposons: phylogenetic misconceptions in the plant world. Mob DNA. 2013;4:9.Bertsch C, Beuve M, Dolja VV, Wirth M, Pelsy F, Herrbach E, et al. Retention of the virus-derived sequences in the nuclear genome of grapevine as a potential pathway to virus resistance. Biol Direct. 2009;4:21.Roossinck MJ. Lifestyle of plant viruses. Philos Trans R Soc B. 2010;365:1899–905.Nibert ML, Ghabrial SA, Maiss E, Lesker T, Vainio EJ, Jiang D, et al. Taxonomic reorganization of family Partitiviridae and other recent progress in partitivirus research. Virus Res. 2014;188:128–41.Koonin EV, Wolf YI, Nagasaki K, Dolja VV. The Big Bang of picorna-like virus evolution antedates the radiation of eukaryotic supergroups. Nat Rev Microbiol. 2008;6:925–39.Poch O, Sauvaget I, Delarue M, Tordo N. Identification of four conserved motifs among the RNA-dependent polymerase encoding elements. EMBO J. 1989;8:3867–74.Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 2010;59:307–21.Letunic I, Bork P. Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 2016;44(W1):W242.Li CX, Shi M, Tian JH, Lin XD, Kang YJ, Chen LJ, et al. Unprecedented genomic diversity of RNA viruses in arthropods reveals the ancestry of negative-sense RNA viruses. eLife. 2015;29:4.Yutin N, Wolf YI, Raoult D, Koonin EV. Eukaryotic large nucleo-cytoplasmic DNA viruses: clusters of orthologous genes and reconstruction of viral genome evolution. Virol J. 2009;6:223.Liu H, Fu Y, Xie J, Cheng J, Ghabrial SA, Li G, et al. Evolutionary genomics of mycovirus-related dsRNA viruses reveals cross-family horizontal gene transfer and evolution of diverse viral lineages. BMC Evol Biol. 2012;12:91.Atsatt PR, Whiteside MD. Novel symbiotic protoplasts formed by endophytic fungi explain their hidden existence, lifestyle switching, and diversity within the plant kingdom. PLoS ONE. 2014;9:e95266.Hom EF, Murray AW. Plant-fungal ecology. Niche engineering demonstrates a latent capacity for fungal-algal mutualism. Science. 2014;345:94–8.Magallón S, Hilu KW, Quandt D. Land plant evolutionary timeline: gene effects are secondary to fossil constraints in relaxed clock estimation of age and substitution rates. Am J Bot. 2013;100:556–73.Johnson MT, Carpenter EJ, Tian Z, Bruskiewich R, Burris JN, Carrigan CT, et al. Evaluating methods for isolating total RNA and predicting the success of sequencing phylogenetically diverse plant transcriptomes. PLoS One. 2012;7:e50226.Kreuze J. siRNA deep sequencing and assembly: piecing together viral infections. In: Gullino ML, Bonants PJM, editors. Detection and diagnostics of plant pathogens. Dordrecht: Springer; 2014. p. 21–38.Chávez Montes RA, de Fátima R-CF, De Paoli E, Accerbi M, Rymarquis LA, Mahalingam G, et al. Sample sequencing of vascular plants demonstrates widespread conservation and divergence of microRNAs. Nat Commun. 2014;5:3722.Marchler-Bauer A, Derbyshire MK, Gonzales NR, Lu S, Chitsaz F, Geer LY, et al. CDD: NCBI’s conserved domain database. Nucl Acids Res. 2015;43:D222–6.Edgar RC. MUSCLE: a multiple sequence alignment method with reduced time and space complexity. BMC Bioinformatics. 2004;5:113.Söding J, Biegert A, Lupas AN. The HHpred interactive server for protein homology detection and structure prediction. Nucleic Acids Res. 2015;33:W244–8.Le SQ, Gascuel O. An improved general amino acid replacement matrix. Mol Biol Evol. 2008;25(7):1307–20.Daydé M, Depardon B, Franc A, Gibrat JF, Guillier R, Karami Y, Sutter F, Taddese B, Chabbert M, Thérond S. E-Biothon: an experimental platform for bioinformatics. 2015 Computer Science and Information Technologies (CSIT) 2015;1-4.Okuno K, Hama T, Takeshita M, Furuya N, Takanami Y. New potyvirus isolated from Cryptotaenia japonica. J Gen Plant Pathol. 2003;69:138–42

    Molecular characterization of a novel ssRNA ourmia-like virus from the rice blast fungus Magnaporthe oryzae

    Get PDF
    In this study we characterize a novel positive and single stranded RNA (ssRNA) mycovirus isolated from the rice field isolate of Magnaporthe oryzae Guy11. The ssRNA contains a single open reading frame (ORF) of 2,373 nucleotides in length and encodes an RNA-dependent RNA polymerase (RdRp) closely related to ourmiaviruses (plant viruses) and ourmia-like mycoviruses. Accordingly, we name this virus Magnaporthe oryzae ourmia-like virus 1 (MOLV1). Although phylogenetic analysis suggests that MOLV1 is closely related to ourmia and ourmia-like viruses, it has some features never reported before within the Ourmiavirus genus. 3' RLM-RACE (RNA ligase-mediated rapid amplification of cDNA ends) and extension poly(A) tests (ePAT) suggest that the MOLV1 genome contains a poly(A) tail whereas the three cytosine and the three guanine residues present in 5' and 3' untranslated regions (UTRs) of ourmia viruses are not observed in the MOLV1 sequence. The discovery of this novel viral genome supports the hypothesis that plant pathogenic fungi may have acquired this type of viruses from their host plants
    corecore