165 research outputs found

    Primase-polymerases are a functionally diverse superfamily of replication and repair enzymes

    Get PDF
    Until relatively recently, DNA primases were viewed simply as a class of proteins that synthesize short RNA primers requisite for the initiation of DNA replication. However, recent studies have shown that this perception of the limited activities associated with these diverse enzymes can no longer be justified. Numerous examples can now be cited demonstrating how the term 'DNA primase' only describes a very narrow subset of these nucleotidyltransferases, with the vast majority fulfilling multifunctional roles from DNA replication to damage tolerance and repair. This article focuses on the archaeo-eukaryotic primase (AEP) superfamily, drawing on recently characterized examples from all domains of life to highlight the functionally diverse pathways in which these enzymes are employed. The broad origins, functionalities and enzymatic capabilities of AEPs emphasizes their previous functional misannotation and supports the necessity for a reclassification of these enzymes under a category called primase-polymerases within the wider functional grouping of polymerases. Importantly, the repositioning of AEPs in this way better recognizes their broader roles in DNA metabolism and encourages the discovery of additional functions for these enzymes, aside from those highlighted here

    DNA synthesis determines the binding mode of the human mitochondrial single-stranded DNA-binding protein

    Get PDF
    [EN] Single-stranded DNA-binding proteins (SSBs) play a key role in genome maintenance, binding and organizing single-stranded DNA (ssDNA) intermediates. Multimeric SSBs, such as the human mitochondrial SSB (HmtSSB), present multiple sites to interact with ssDNA, which has been shown in vitro to enable them to bind a variable number of single-stranded nucleotides depending on the salt and protein concentration. It has long been suggested that different binding modes might be used selectively for different functions. To study this possibility, we used optical tweezers to determine and compare the structure and energetics of long, individual HmtSSB¿DNA complexes assembled on preformed ssDNA and on ssDNA generated gradually during `in situ¿ DNA synthesis. We show that HmtSSB binds to preformed ss-DNA in two major modes, depending on salt and protein concentration. However, when protein binding was coupled to strand-displacement DNA synthesis, only one of the two binding modes was observed under all experimental conditions. Our results reveal a key role for the gradual generation of ssDNA in modulating the binding mode of a multimeric SSB protein and consequently, in generating the appropriate nucleoprotein structure for DNA synthetic reactions required for genome maintenance.We are grateful to Prof. M. Salas laboratory (CBMSO-CSIC) for generously providing the Phi29 DNA polymerase and to Juan P. García Villaluenga (UCM) for useful discussions. Spanish Ministry of Economy and Competitiveness [MAT2015-71806-R to J.R.A-G, FIS2010-17440, FIS2015-67765-R to F.J.C., BFU2012-31825, BFU2015-63714-R to B.I.]; Spanish Ministry of Education, Culture and Sport [FPU13/02934 to J.J., FPU13/02826 to E.B-H.]; National Institutes of Health [GM45925 to L.S.K.]; University of Tampere (to G.L.C.); Programa de Financiacion Universidad Complutense de Madrid-Santander Universidades [CT45/15-CT46/15 to F.C.]. Funding for open access charge: Spanish Ministry of Economy and Competitiveness [BFU2015-63714-R].Morin, J.; Cerrón, F.; Jarillo, J.; Beltran-Heredia, E.; Ciesielski, G.; Arias-Gonzalez, JR.; Kaguni, L.... (2017). DNA synthesis determines the binding mode of the human mitochondrial single-stranded DNA-binding protein. Nucleic Acids Research. 45(12):7237-7248. https://doi.org/10.1093/nar/gkx395S723772484512Shereda, R. D., Kozlov, A. G., Lohman, T. M., Cox, M. M., & Keck, J. L. (2008). SSB as an Organizer/Mobilizer of Genome Maintenance Complexes. Critical Reviews in Biochemistry and Molecular Biology, 43(5), 289-318. doi:10.1080/10409230802341296Flynn, R. L., & Zou, L. (2010). Oligonucleotide/oligosaccharide-binding fold proteins: a growing family of genome guardians. Critical Reviews in Biochemistry and Molecular Biology, 45(4), 266-275. doi:10.3109/10409238.2010.488216Murzin, A. G. (1993). OB(oligonucleotide/oligosaccharide binding)-fold: common structural and functional solution for non-homologous sequences. The EMBO Journal, 12(3), 861-867. doi:10.1002/j.1460-2075.1993.tb05726.xKozlov, A. G., Weiland, E., Mittal, A., Waldman, V., Antony, E., Fazio, N., … Lohman, T. M. (2015). Intrinsically Disordered C-Terminal Tails of E. coli Single-Stranded DNA Binding Protein Regulate Cooperative Binding to Single-Stranded DNA. Journal of Molecular Biology, 427(4), 763-774. doi:10.1016/j.jmb.2014.12.020Kuznetsov, S. V., Kozlov, A. G., Lohman, T. M., & Ansari, A. (2006). Microsecond Dynamics of Protein–DNA Interactions: Direct Observation of the Wrapping/Unwrapping Kinetics of Single-stranded DNA around the E.coli SSB Tetramer. Journal of Molecular Biology, 359(1), 55-65. doi:10.1016/j.jmb.2006.02.070Lohman, T. M., & Ferrari, M. E. (1994). Escherichia Coli Single-Stranded DNA-Binding Protein: Multiple DNA-Binding Modes and Cooperativities. Annual Review of Biochemistry, 63(1), 527-570. doi:10.1146/annurev.bi.63.070194.002523Maier, D., Farr, C. L., Poeck, B., Alahari, A., Vogel, M., Fischer, S., … Schneuwly, S. (2001). Mitochondrial Single-stranded DNA-binding Protein Is Required for Mitochondrial DNA Replication and Development in Drosophila melanogaster. Molecular Biology of the Cell, 12(4), 821-830. doi:10.1091/mbc.12.4.821Ruhanen, H., Borrie, S., Szabadkai, G., Tyynismaa, H., Jones, A. W. E., Kang, D., … Yasukawa, T. (2010). Mitochondrial single-stranded DNA binding protein is required for maintenance of mitochondrial DNA and 7S DNA but is not required for mitochondrial nucleoid organisation. Biochimica et Biophysica Acta (BBA) - Molecular Cell Research, 1803(8), 931-939. doi:10.1016/j.bbamcr.2010.04.008Farr, C. L., Matsushima, Y., Lagina, A. T., Luo, N., & Kaguni, L. S. (2004). Physiological and Biochemical Defects in Functional Interactions of Mitochondrial DNA Polymerase and DNA-binding Mutants of Single-stranded DNA-binding Protein. Journal of Biological Chemistry, 279(17), 17047-17053. doi:10.1074/jbc.m400283200Van Tuyle, G. C., & Pavco, P. A. (1985). The rat liver mitochondrial DNA-protein complex: displaced single strands of replicative intermediates are protein coated. The Journal of Cell Biology, 100(1), 251-257. doi:10.1083/jcb.100.1.251Clayton, D. A. (1982). Replication of animal mitochondrial DNA. Cell, 28(4), 693-705. doi:10.1016/0092-8674(82)90049-6Farr, C. L., Wang, Y., & Kaguni, L. S. (1999). Functional Interactions of Mitochondrial DNA Polymerase and Single-stranded DNA-binding Protein. Journal of Biological Chemistry, 274(21), 14779-14785. doi:10.1074/jbc.274.21.14779Korhonen, J. A., Gaspari, M., & Falkenberg, M. (2003). TWINKLE Has 5′ → 3′ DNA Helicase Activity and Is Specifically Stimulated by Mitochondrial Single-stranded DNA-binding Protein. Journal of Biological Chemistry, 278(49), 48627-48632. doi:10.1074/jbc.m306981200Miralles Fusté, J., Shi, Y., Wanrooij, S., Zhu, X., Jemt, E., Persson, Ö., … Falkenberg, M. (2014). In Vivo Occupancy of Mitochondrial Single-Stranded DNA Binding Protein Supports the Strand Displacement Mode of DNA Replication. PLoS Genetics, 10(12), e1004832. doi:10.1371/journal.pgen.1004832Oliveira, M. T., & Kaguni, L. S. (2011). Reduced Stimulation of Recombinant DNA Polymerase γ and Mitochondrial DNA (mtDNA) Helicase by Variants of Mitochondrial Single-stranded DNA-binding Protein (mtSSB) Correlates with Defects in mtDNA Replication in Animal Cells. Journal of Biological Chemistry, 286(47), 40649-40658. doi:10.1074/jbc.m111.289983Williams, A. J., & Kaguni, L. S. (1995). Stimulation ofDrosophilaMitochondrial DNA Polymerase by Single-stranded DNA-binding Protein. Journal of Biological Chemistry, 270(2), 860-865. doi:10.1074/jbc.270.2.860Bogenhagen, D. F., Wang, Y., Shen, E. L., & Kobayashi, R. (2003). Protein Components of Mitochondrial DNA Nucleoids in Higher Eukaryotes. Molecular & Cellular Proteomics, 2(11), 1205-1216. doi:10.1074/mcp.m300035-mcp200BARAT-GUERIDE, M., DUFRESNE, C., & RICKWOOD, D. (1989). Effect of DNA conformation on the transcription of mitochondrial DNA. European Journal of Biochemistry, 183(2), 297-302. doi:10.1111/j.1432-1033.1989.tb14928.xYang, C., Curth, U., Urbanke, C., & Kang, C. (1997). Crystal structure of human mitochondrial single-stranded DNA binding protein at 2.4 Å resolution. Nature Structural Biology, 4(2), 153-157. doi:10.1038/nsb0297-153Raghunathan, S., Ricard, C. S., Lohman, T. M., & Waksman, G. (1997). Crystal structure of the homo-tetrameric DNA binding domain of Escherichia coli single-stranded DNA-binding protein determined by multiwavelength x-ray diffraction on the selenomethionyl protein at 2.9-A resolution. Proceedings of the National Academy of Sciences, 94(13), 6652-6657. doi:10.1073/pnas.94.13.6652CURTH, U., URBANKE, C., GREIPEL, J., GERBERDING, H., TIRANTI, V., & ZEVIANI, M. (1994). Single-stranded-DNA-binding proteins from human mitochondria and Escherichia coli have analogous physicochemical properties. European Journal of Biochemistry, 221(1), 435-443. doi:10.1111/j.1432-1033.1994.tb18756.xOverman, L. B., & Lohman, T. M. (1994). Linkage of pH, Anion and Cation Effects in Protein-Nucleic Acid Equilibria. Journal of Molecular Biology, 236(1), 165-178. doi:10.1006/jmbi.1994.1126Bhattacharyya, B., George, N. P., Thurmes, T. M., Zhou, R., Jani, N., Wessel, S. R., … Keck, J. L. (2013). Structural mechanisms of PriA-mediated DNA replication restart. Proceedings of the National Academy of Sciences, 111(4), 1373-1378. doi:10.1073/pnas.1318001111Carlini, L. E., Porter, R. D., Curth, U., & Urbanke, C. (1993). Viability and preliminary in vivo characterization of site directed mutants of Escherichia coli single-stranded DNA-binding protein. Molecular Microbiology, 10(5), 1067-1075. doi:10.1111/j.1365-2958.1993.tb00977.xGriffith, J. D., Harris, L. D., & Register, J. (1984). Visualization of SSB-ssDNA Complexes Active in the Assembly of Stable RecA-DNA Filaments. Cold Spring Harbor Symposia on Quantitative Biology, 49(0), 553-559. doi:10.1101/sqb.1984.049.01.062Morrical, S. W., & Cox, M. M. (1990). Stabilization of recA protein-ssDNA complexes by the single-stranded DNA binding protein of Escherichia coli. Biochemistry, 29(3), 837-843. doi:10.1021/bi00455a034Muniyappa, K., Williams, K., Chase, J. W., & Radding, C. M. (1990). Active nucleoprotein filaments of single-stranded binding protein and recA protein on single-stranded DNA have a regular repeating structure. Nucleic Acids Research, 18(13), 3967-3973. doi:10.1093/nar/18.13.3967Wessel, S. R., Marceau, A. H., Massoni, S. C., Zhou, R., Ha, T., Sandler, S. J., & Keck, J. L. (2013). PriC-mediated DNA Replication Restart Requires PriC Complex Formation with the Single-stranded DNA-binding Protein. Journal of Biological Chemistry, 288(24), 17569-17578. doi:10.1074/jbc.m113.478156Bell, J. C., Liu, B., & Kowalczykowski, S. C. (2015). Imaging and energetics of single SSB-ssDNA molecules reveal intramolecular condensation and insight into RecOR function. eLife, 4. doi:10.7554/elife.08646Suksombat, S., Khafizov, R., Kozlov, A. G., Lohman, T. M., & Chemla, Y. R. (2015). Structural dynamics of E. coli single-stranded DNA binding protein reveal DNA wrapping and unwrapping pathways. eLife, 4. doi:10.7554/elife.08193Zhou, R., Kozlov, A. G., Roy, R., Zhang, J., Korolev, S., Lohman, T. M., & Ha, T. (2011). SSB Functions as a Sliding Platform that Migrates on DNA via Reptation. Cell, 146(2), 222-232. doi:10.1016/j.cell.2011.06.036Pant, K., Karpel, R. L., Rouzina, I., & Williams, M. C. (2004). Mechanical Measurement of Single-molecule Binding Rates: Kinetics of DNA Helix-destabilization by T4 Gene 32 Protein. Journal of Molecular Biology, 336(4), 851-870. doi:10.1016/j.jmb.2003.12.025Pant, K., Karpel, R. L., Rouzina, I., & Williams, M. C. (2005). Salt Dependent Binding of T4 Gene 32 Protein to Single and Double-stranded DNA: Single Molecule Force Spectroscopy Measurements. Journal of Molecular Biology, 349(2), 317-330. doi:10.1016/j.jmb.2005.03.065Robberson, D. L., & Clayton, D. A. (1972). Replication of Mitochondrial DNA in Mouse L Cells and Their Thymidine Kinase- Derivatives: Displacement Replication on a Covalently-Closed Circular Template. Proceedings of the National Academy of Sciences, 69(12), 3810-3814. doi:10.1073/pnas.69.12.3810Ciesielski, G. L., Bermek, O., Rosado-Ruiz, F. A., Hovde, S. L., Neitzke, O. J., Griffith, J. D., & Kaguni, L. S. (2015). Mitochondrial Single-stranded DNA-binding Proteins Stimulate the Activity of DNA Polymerase γ by Organization of the Template DNA. Journal of Biological Chemistry, 290(48), 28697-28707. doi:10.1074/jbc.m115.673707Lázaro, J. M., Blanco, L., & Salas, M. (1995). [5] Purification of bacteriophage φ29 DNA polymerase. DNA Replication, 42-49. doi:10.1016/0076-6879(95)62007-9Ibarra, B., Chemla, Y. R., Plyasunov, S., Smith, S. B., Lázaro, J. M., Salas, M., & Bustamante, C. (2009). Proofreading dynamics of a processive DNA polymerase. The EMBO Journal, 28(18), 2794-2802. doi:10.1038/emboj.2009.219Morin, J. A., Cao, F. J., Lazaro, J. M., Arias-Gonzalez, J. R., Valpuesta, J. M., Carrascosa, J. L., … Ibarra, B. (2012). Active DNA unwinding dynamics during processive DNA replication. Proceedings of the National Academy of Sciences, 109(21), 8115-8120. doi:10.1073/pnas.1204759109Smith, S. B., Cui, Y., & Bustamante, C. (2003). [7] Optical-trap force transducer that operates by direct measurement of light momentum. Biophotonics, Part B, 134-162. doi:10.1016/s0076-6879(03)61009-8Bosco, A., Camunas-Soler, J., & Ritort, F. (2013). Elastic properties and secondary structure formation of single-stranded DNA at monovalent and divalent salt conditions. Nucleic Acids Research, 42(3), 2064-2074. doi:10.1093/nar/gkt1089Smith, S., Finzi, L., & Bustamante, C. (1992). Direct mechanical measurements of the elasticity of single DNA molecules by using magnetic beads. Science, 258(5085), 1122-1126. doi:10.1126/science.1439819Longley, M. J., Smith, L. A., & Copeland, W. C. (2009). Preparation of Human Mitochondrial Single-Stranded DNA-Binding Protein. Mitochondrial DNA, 73-85. doi:10.1007/978-1-59745-521-3_5Li, K., & Williams, R. S. (1997). Tetramerization and Single-stranded DNA Binding Properties of Native and Mutated Forms of Murine Mitochondrial Single-stranded DNA-binding Proteins. Journal of Biological Chemistry, 272(13), 8686-8694. doi:10.1074/jbc.272.13.8686Jarillo, J., Morín, J. A., Beltrán-Heredia, E., Villaluenga, J. P. G., Ibarra, B., & Cao, F. J. (2017). Mechanics, thermodynamics, and kinetics of ligand binding to biopolymers. PLOS ONE, 12(4), e0174830. doi:10.1371/journal.pone.0174830Bujalowski, W., & Lohman, T. M. (1986). Escherichia coli single-strand binding protein forms multiple, distinct complexes with single-stranded DNA. Biochemistry, 25(24), 7799-7802. doi:10.1021/bi00372a003Thömmes, P., Farr, C. L., Marton, R. F., Kaguni, L. S., & Cotterill, S. (1995). Mitochondrial Single-stranded DNA-binding Protein fromDrosophilaEmbryos. Journal of Biological Chemistry, 270(36), 21137-21143. doi:10.1074/jbc.270.36.21137Rodriguez, I., Lazaro, J. M., Blanco, L., Kamtekar, S., Berman, A. J., Wang, J., … de Vega, M. (2005). A specific subdomain in  29 DNA polymerase confers both processivity and strand-displacement capacity. Proceedings of the National Academy of Sciences, 102(18), 6407-6412. doi:10.1073/pnas.0500597102Kamtekar, S., Berman, A. J., Wang, J., Lázaro, J. M., de Vega, M., Blanco, L., … Steitz, T. A. (2004). Insights into Strand Displacement and Processivity from the Crystal Structure of the Protein-Primed DNA Polymerase of Bacteriophage φ29. Molecular Cell, 16(4), 609-618. doi:10.1016/j.molcel.2004.10.019Chrysogelos, S., & Griffith, J. (1982). Escherichia coli single-strand binding protein organizes single-stranded DNA in nucleosome-like units. Proceedings of the National Academy of Sciences, 79(19), 5803-5807. doi:10.1073/pnas.79.19.5803Hamon, L., Pastre, D., Dupaigne, P., Breton, C. L., Cam, E. L., & Pietrement, O. (2007). High-resolution AFM imaging of single-stranded DNA-binding (SSB) protein--DNA complexes. Nucleic Acids Research, 35(8), e58-e58. doi:10.1093/nar/gkm147Takamatsu, C., Umeda, S., Ohsato, T., Ohno, T., Abe, Y., Fukuoh, A., … Kang, D. (2002). Regulation of mitochondrial D‐loops by transcription factor A and single‐stranded DNA‐binding protein. EMBO reports, 3(5), 451-456. doi:10.1093/embo-reports/kvf099Wang, Y., & Bogenhagen, D. F. (2006). Human Mitochondrial DNA Nucleoids Are Linked to Protein Folding Machinery and Metabolic Enzymes at the Mitochondrial Inner Membrane. Journal of Biological Chemistry, 281(35), 25791-25802. doi:10.1074/jbc.m604501200Brown, T. A. (2005). Replication of mitochondrial DNA occurs by strand displacement with alternative light-strand origins, not via a strand-coupled mechanism. Genes & Development, 19(20), 2466-2476. doi:10.1101/gad.135210

    Modeling Pathogenic Mutations of Human Twinkle in Drosophila Suggests an Apoptosis Role in Response to Mitochondrial Defects

    Get PDF
    The human gene C10orf2 encodes the mitochondrial replicative DNA helicase Twinkle, mutations of which are responsible for a significant fraction of cases of autosomal dominant progressive external ophthalmoplegia (adPEO), a human mitochondrial disease caused by defects in intergenomic communication. We report the analysis of orthologous mutations in the Drosophila melanogaster mitochondrial DNA (mtDNA) helicase gene, d-mtDNA helicase. Increased expression of wild type d-mtDNA helicase using the UAS-GAL4 system leads to an increase in mtDNA copy number throughout adult life without any noteworthy phenotype, whereas overexpression of d-mtDNA helicase containing the K388A mutation in the helicase active site results in a severe depletion of mtDNA and a lethal phenotype. Overexpression of two d-mtDNA helicase variants equivalent to two human adPEO mutations shows differential effects. The A442P mutation exhibits a dominant negative effect similar to that of the active site mutant. In contrast, overexpression of d-mtDNA helicase containing the W441C mutation results in a slight decrease in mtDNA copy number during the third instar larval stage, and a moderate decrease in life span in the adult population. Overexpression of d-mtDNA helicase containing either the K388A or A442P mutations causes a mitochondrial oxidative phosphorylation (OXPHOS) defect that significantly reduces cell proliferation. The mitochondrial impairment caused by these mutations promotes apoptosis, arguing that mitochondria regulate programmed cell death in Drosophila. Our study of d-mtDNA helicase overexpression provides a tractable Drosophila model for understanding the cellular and molecular effects of human adPEO mutations

    Mitochondrial Single-stranded DNA-binding Proteins Stimulate the Activity of DNA Polymerase γ by Organization of the Template DNA

    Get PDF
    The activity of the mitochondrial replicase, DNA polymerase γ (Pol γ) is stimulated by another key component of the mitochondrial replisome, the mitochondrial single-stranded DNA-binding protein (mtSSB). We have performed a comparative analysis of the human and Drosophila Pols γ with their cognate mtSSBs, evaluating their functional relationships using a combined approach of biochemical assays and electron microscopy. We found that increasing concentrations of both mtSSBs led to the elimination of template secondary structure and gradual opening of the template DNA, through a series of visually similar template species. The stimulatory effect of mtSSB on Pol γ on these ssDNA templates is not species-specific. We observed that human mtSSB can be substituted by its Drosophila homologue, and vice versa, finding that a lower concentration of insect mtSSB promotes efficient stimulation of either Pol. Notably, distinct phases of the stimulation by both mtSSBs are distinguishable, and they are characterized by a similar organization of the template DNA for both Pols γ. We conclude that organization of the template DNA is the major factor contributing to the stimulation of Pol γ activity. Additionally, we observed that human Pol γ preferentially utilizes compacted templates, whereas the insect enzyme achieves its maximal activity on open templates, emphasizing the relative importance of template DNA organization in modulating Pol γ activity and the variation among systems

    Disruption of Mitochondrial DNA Replication in Drosophila Increases Mitochondrial Fast Axonal Transport In Vivo

    Get PDF
    Mutations in mitochondrial DNA polymerase (pol γ) cause several progressive human diseases including Parkinson's disease, Alper's syndrome, and progressive external ophthalmoplegia. At the cellular level, disruption of pol γ leads to depletion of mtDNA, disrupts the mitochondrial respiratory chain, and increases susceptibility to oxidative stress. Although recent studies have intensified focus on the role of mtDNA in neuronal diseases, the changes that take place in mitochondrial biogenesis and mitochondrial axonal transport when mtDNA replication is disrupted are unknown. Using high-speed confocal microscopy, electron microscopy and biochemical approaches, we report that mutations in pol γ deplete mtDNA levels and lead to an increase in mitochondrial density in Drosophila proximal nerves and muscles, without a noticeable increase in mitochondrial fragmentation. Furthermore, there is a rise in flux of bidirectional mitochondrial axonal transport, albeit with slower kinesin-based anterograde transport. In contrast, flux of synaptic vesicle precursors was modestly decreased in pol γ−α mutants. Our data indicate that disruption of mtDNA replication does not hinder mitochondrial biogenesis, increases mitochondrial axonal transport, and raises the question of whether high levels of circulating mtDNA-deficient mitochondria are beneficial or deleterious in mtDNA diseases

    C-Terminal Extension of the Yeast Mitochondrial DNA Polymerase Determines the Balance between Synthesis and Degradation

    Get PDF
    Saccharomyces cerevisiae mitochondrial DNA polymerase (Mip1) contains a C-terminal extension (CTE) of 279 amino acid residues. The CTE is required for mitochondrial DNA maintenance in yeast but is absent in higher eukaryotes. Here we use recombinant Mip1 C-terminal deletion mutants to investigate functional importance of the CTE. We show that partial removal of the CTE in Mip1Δ216 results in strong preference for exonucleolytic degradation rather than DNA polymerization. This disbalance in exonuclease and polymerase activities is prominent at suboptimal dNTP concentrations and in the absence of correctly pairing nucleotide. Mip1Δ216 also displays reduced ability to synthesize DNA through double-stranded regions. Full removal of the CTE in Mip1Δ279 results in complete loss of Mip1 polymerase activity, however the mutant retains its exonuclease activity. These results allow us to propose that CTE functions as a part of Mip1 polymerase domain that stabilizes the substrate primer end at the polymerase active site, and is therefore required for efficient mitochondrial DNA replication in vivo

    POLG1 p.R722H mutation associated with multiple mtDNA deletions and a neurological phenotype

    Get PDF
    <p>Abstract</p> <p>Background</p> <p>The c.2447G>A (p.R722H) mutation in the gene <it>POLG1 </it>of the catalytic subunit of human mitochondrial polymerase gamma has been previously found in a few occasions but its pathogenicity has remained uncertain. We set out to ascertain its contribution to neuromuscular disease.</p> <p>Methods</p> <p>Probands from two families with probable mitochondrial disease were examined clinically, muscle and buccal epithelial DNA were analyzed for mtDNA deletions, and the <it>POLG1, POLG2, ANT1 </it>and <it>Twinkle </it>genes were sequenced.</p> <p>Results</p> <p>An adult proband presented with progressive external ophthalmoplegia, sensorineural hearing impairment, diabetes mellitus, dysphagia, a limb myopathy and dementia. Brain MRI showed central and cortical atrophy, and <sup>18</sup>F-deoxyglucose PET revealed reduced glucose uptake. Histochemical analysis of muscle disclosed ragged red fibers and cytochrome c oxidase-negative fibers. Electron microscopy showed subsarcolemmal aggregates of morphologically normal mitochondria. Multiple mtDNA deletions were found in the muscle, and sequencing of the <it>POLG1 </it>gene revealed a homozygous c.2447G>A (p.R722H) mutation. His two siblings were also homozygous with respect to the p.R722H mutation and presented with dementia and sensorineural hearing impairment. In another family the p.R722H mutation was found as compound heterozygosity with the common p.W748S mutation in two siblings with mental retardation, ptosis, epilepsy and psychiatric symptoms. The estimated carrier frequency of the p.R722H mutation was 1:135 in the Finnish population. No mutations in <it>POLG2</it>, <it>ANT1 </it>and <it>Twinkle </it>genes were found. Analysis of the POLG1 sequence by homology modeling supported the notion that the p.R722H mutation is pathogenic.</p> <p>Conclusions</p> <p>The recessive c.2447G>A (p.R722H) mutation in the linker region of the <it>POLG1 </it>gene is pathogenic for multiple mtDNA deletions in muscle and is associated with a late-onset neurological phenotype as a homozygous state. The onset of the disease can be earlier in compound heterozygotes.</p

    Association of the Chromosome Replication Initiator DnaA with the Escherichia coli Inner Membrane In Vivo: Quantity and Mode of Binding

    Get PDF
    DnaA initiates chromosome replication in most known bacteria and its activity is controlled so that this event occurs only once every cell division cycle. ATP in the active ATP-DnaA is hydrolyzed after initiation and the resulting ADP is replaced with ATP on the verge of the next initiation. Two putative recycling mechanisms depend on the binding of DnaA either to the membrane or to specific chromosomal sites, promoting nucleotide dissociation. While there is no doubt that DnaA interacts with artificial membranes in vitro, it is still controversial as to whether it binds the cytoplasmic membrane in vivo. In this work we looked for DnaA-membrane interaction in E. coli cells by employing cell fractionation with both native and fluorescent DnaA hybrids. We show that about 10% of cellular DnaA is reproducibly membrane-associated. This small fraction might be physiologically significant and represent the free DnaA available for initiation, rather than the vast majority bound to the datA reservoir. Using the combination of mCherry with a variety of DnaA fragments, we demonstrate that the membrane binding function is delocalized on the surface of the protein’s domain III, rather than confined to a particular sequence. We propose a new binding-bending mechanism to explain the membrane-induced nucleotide release from DnaA. This mechanism would be fundamental to the initiation of replication

    Classification of Plant Associated Bacteria Using RIF, a Computationally Derived DNA Marker

    Get PDF
    A DNA marker that distinguishes plant associated bacteria at the species level and below was derived by comparing six sequenced genomes of Xanthomonas, a genus that contains many important phytopathogens. This DNA marker comprises a portion of the dnaA replication initiation factor (RIF). Unlike the rRNA genes, dnaA is a single copy gene in the vast majority of sequenced bacterial genomes, and amplification of RIF requires genus-specific primers. In silico analysis revealed that RIF has equal or greater ability to differentiate closely related species of Xanthomonas than the widely used ribosomal intergenic spacer region (ITS). Furthermore, in a set of 263 Xanthomonas, Ralstonia and Clavibacter strains, the RIF marker was directly sequenced in both directions with a success rate approximately 16% higher than that for ITS. RIF frameworks for Xanthomonas, Ralstonia and Clavibacter were constructed using 682 reference strains representing different species, subspecies, pathovars, races, hosts and geographic regions, and contain a total of 109 different RIF sequences. RIF sequences showed subspecific groupings but did not place strains of X. campestris or X. axonopodis into currently named pathovars nor R. solanacearum strains into their respective races, confirming previous conclusions that pathovar and race designations do not necessarily reflect genetic relationships. The RIF marker also was sequenced for 24 reference strains from three genera in the Enterobacteriaceae: Pectobacterium, Pantoea and Dickeya. RIF sequences of 70 previously uncharacterized strains of Ralstonia, Clavibacter, Pectobacterium and Dickeya matched, or were similar to, those of known reference strains, illustrating the utility of the frameworks to classify bacteria below the species level and rapidly match unknown isolates to reference strains. The RIF sequence frameworks are available at the online RIF database, RIFdb, and can be queried for diagnostic purposes with RIF sequences obtained from unknown strains in both chromatogram and FASTA format
    corecore