518 research outputs found

    Helicase Loading at Chromosomal Origins of Replication

    Get PDF
    Loading of the replicative DNA helicase at origins of replication is of central importance in DNA replication. As the first of the replication fork proteins assemble at chromosomal origins of replication, the loaded helicase is required for the recruitment of the rest of the replication machinery. In this work, we review the current knowledge of helicase loading at Escherichia coli and eukaryotic origins of replication. In each case, this process requires both an origin recognition protein as well as one or more additional proteins. Comparison of these events shows intriguing similarities that suggest a similar underlying mechanism, as well as critical differences that likely reflect the distinct processes that regulate helicase loading in bacterial and eukaryotic cells.Howard Hughes Medical InstituteNational Institutes of Health (U.S.) (Grant RO1 GM052339

    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

    Lethal Interaction of Nuclear and Mitochondrial Genotypes in Drosophila melanogaster

    Get PDF
    Drosophila melanogaster, like most animal species, displays considerable genetic variation in both nuclear and mitochondrial DNA (mtDNA). Here we tested whether any of four natural mtDNA variants was able to modify the effect of the phenotypically mild, nuclear tko(25t) mutation, affecting mitochondrial protein synthesis. When combined with tko(25t), the mtDNA from wild strain KSA2 produced pupal lethality, accompanied by the presence of melanotic nodules in L3 larvae. KSA2 mtDNA, which carries a substitution at a conserved residue of cytochrome b that is predicted to be involved in subunit interactions within respiratory complex III, conferred drastically decreased respiratory capacity and complex III activity in the tko(25t) but not a wild-type nuclear background. The complex III inhibitor antimycin A was able to phenocopy effects of the tko(25t) mutation in the KSA2 mtDNA background. This is the first report of a lethal, nuclear-mitochondrial interaction within a metazoan species, representing a paradigm for understanding genetic interactions between nuclear and mitochondrial genotype relevant to human health and disease.Peer reviewe

    A cytoplasmic suppressor of a nuclear mutation affecting mitochondrial functions in Drosophila

    Get PDF
    Phenotypes relevant to oxidative phosphorylation (OXPHOS) in eukaryotes are jointly determined by nuclear and mitochondrial DNA (mtDNA). Thus, in humans, the variable clinical presentations of mitochondrial disease patients bearing the same primary mutation, whether in nuclear or mitochondrial DNA, have been attributed to putative genetic determinants carried in the other genome, though their identity and the molecular mechanism(s) by which they might act remain elusive. Here we demonstrate cytoplasmic suppression of the mitochondrial disease-like phenotype of the Drosophila melanogaster nuclear mutant tko25t, which includes developmental delay, seizure sensitivity, and defective male courtship. The tko25t strain carries a mutation in a mitoribosomal protein gene, causing OXPHOS deficiency due to defective intramitochondrial protein synthesis. Phenotypic suppression was associated with increased mtDNA copy number and increased mitochondrial biogenesis, as measured by the expression levels of porin voltage dependent anion channel and Spargel (PGC1α). Ubiquitous overexpression of Spargel in tko25t flies phenocopied the suppressor, identifying it as a key mechanistic target thereof. Suppressor-strain mtDNAs differed from related nonsuppressor strain mtDNAs by several coding-region polymorphisms and by length and sequence variation in the noncoding region (NCR), in which the origin of mtDNA replication is located. Cytoplasm from four of five originally Wolbachia-infected strains showed the same suppressor effect, whereas that from neither of two uninfected strains did so, suggesting that the stress of chronic Wolbachia infection may provide evolutionary selection for improved mitochondrial fitness under metabolic stress. Our findings provide a paradigm for understanding the role of mtDNA genotype in human disease

    Dynamic assembly of Hda and the sliding clamp in the regulation of replication licensing

    Get PDF
    Regulatory inactivation of DnaA (RIDA) is one of the major regulatory mechanisms of prokaryotic replication licensing. In RIDA, the Hda-sliding clamp complex loaded onto DNA directly interacts with adenosine triphosphate (ATP)-bound DnaA and stimulates the hydrolysis of ATP to inactivate DnaA. A prediction is that the activity of Hda is tightly controlled to ensure that replication initiation occurs only once per cell cycle. Here, we determined the crystal structure of the Hda-�� clamp complex. This complex contains two pairs of Hda dimers sandwiched between two �� clamp rings to form an octamer that is stabilized by three discrete interfaces. Two separate surfaces of Hda make contact with the �� clamp, which is essential for Hda function in RIDA. The third interface between Hda monomers occludes the active site arginine finger, blocking its access to DnaA. Taken together, our structural and mutational analyses of the Hda-�� clamp complex indicate that the interaction of the �� clamp with Hda controls the ability of Hda to interact with DnaA. In the octameric Hda-�� clamp complex, the inability of Hda to interact with DnaA is a novel mechanism that may regulate Hda function. ? The Author(s) 2017.113Ysciescopu

    Structural rearrangements in the mitochondrial genome of Drosophila melanogaster induced by elevated levels of the replicative DNA helicase

    Get PDF
    Pathological conditions impairing functions of mitochondria often lead to compensatory upregulation of the mitochondrial DNA (mtDNA) replisome machinery, and the replicative DNA helicase appears to be a key factor in regulating mtDNA copy number. Moreover, mtDNA helicase mutations have been associated with structural rearrangements of themitochondrial genome. To evaluate the effects of elevated levels of the mtDNA helicase on the integrity and replication of the mitochondrial genome, we overexpressed the helicase in Drosophila melanogaster Schneider cells and analyzed the mtDNA by two-dimensional neutral agarose gel electrophoresis and electron microscopy. We found that elevation of mtDNA helicase levels increases the quantity of replication intermediates and alleviates pausing at the replication slow zones. Though we did not observe a concomitant alteration in mtDNA copy number, we observed deletions specific to the segment of repeated elements in the immediate vicinity of the origin of replication, and an accumulation of species characteristic of replication fork stalling. We also found elevated levels of RNA that are retained in the replication intermediates. Together, our results suggest that upregulation of mtDNA helicase promotes the process of mtDNA replication but also results in genome destabilization.Peer reviewe

    A cytoplasmic suppressor of a nuclear mutation affecting mitochondrial functions in Drosophila

    Get PDF
    Phenotypes relevant to oxidative phosphorylation (OXPHOS) in eukaryotes are jointly determined by nuclear and mitochondrial DNA (mtDNA). Thus, in humans, the variable clinical presentations of mitochondrial disease patients bearing the same primary mutation, whether in nuclear or mitochondrial DNA, have been attributed to putative genetic determinants carried in the other genome, though their identity and the molecular mechanism(s) by which they might act remain elusive. Here we demonstrate cytoplasmic suppression of the mitochondrial disease-like phenotype of the Drosophila melanogaster nuclear mutant tko25t, which includes developmental delay, seizure sensitivity, and defective male courtship. The tko25t strain carries a mutation in a mitoribosomal protein gene, causing OXPHOS deficiency due to defective intramitochondrial protein synthesis. Phenotypic suppression was associated with increased mtDNA copy number and increased mitochondrial biogenesis, as measured by the expression levels of porin voltage dependent anion channel and Spargel (PGC1α). Ubiquitous overexpression of Spargel in tko25t flies phenocopied the suppressor, identifying it as a key mechanistic target thereof. Suppressor-strain mtDNAs differed from related nonsuppressor strain mtDNAs by several coding-region polymorphisms and by length and sequence variation in the noncoding region (NCR), in which the origin of mtDNA replication is located. Cytoplasm from four of five originally Wolbachia-infected strains showed the same suppressor effect, whereas that from neither of two uninfected strains did so, suggesting that the stress of chronic Wolbachia infection may provide evolutionary selection for improved mitochondrial fitness under metabolic stress. Our findings provide a paradigm for understanding the role of mtDNA genotype in human disease

    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

    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
    corecore