Replication of G Quadruplex DNA
Abstract
:1. The Nature of DNA Secondary Structures
2. A Technical Aside: Monitoring the Response of the Replisome to Sites of Secondary Structure Formation In Vivo
3. G Quadruplex (G4)-Forming Potential in the Genome Is Associated with Genetic and Epigenetic Instability
4. Evidence That G4s Can Interfere with Both Lagging and Leading Strand Replication
4.1. Lagging Strand G4s
4.2. Leading Strand G4s
5. When Do G4s Form and How Do They Interact with the Replisome?
5.1. How Could Transcription Promote G4 Formation?
5.2. G4 Formation and the Replisome.
5.3. Pre-Formed G4s Encountered by the Replisome
5.4. Formation of G4s within the Replisome
6. Countering G4 Formation during Replication
7. Helicases Involved in G4 Processing
7.1. Superfamily 1: DNA2 and PIF1
7.1.1. DNA2
7.1.2. PIF1
7.2. The Superfamily 2 RecQ-Like Helicases: BLM and WRN
7.2.1. BLM
7.2.2. WRN
7.3. The Superfamily 2 Fe-S Helicases: RTEL1, DDX11, FANCJ and XPD
7.3.1. FANCJ (Also Known as BRIP1 and BACH1)
7.3.2. DDX11 (Also Known as CHLR1)
7.3.3. RTEL1
7.3.4. XPD/ERCC2
7.4. The Superfamily 2 DEAD-Box Helicases: DHX36 (Also Known as RHAU/G4-Resolvase 1) and DHX9
7.4.1. DHX36
7.4.2. DHX9 (Also Known as Nuclear DNA helicase II/RNA Helicase A)
8. Coda: Principles Governing the Recruitment of G4 Helicases
8.1. Selection
8.2. Collaboration
Funding
Acknowledgments
Conflicts of Interest
References
- Mirkin, S.M. DNA structures, repeat expansions and human hereditary disorders. Curr. Opin. Struct. Biol. 2006, 16, 351–358. [Google Scholar] [CrossRef] [PubMed]
- Wickramasinghe, C.M.; Arzouk, H.; Frey, A.; Maiter, A.; Sale, J.E. Contributions of the specialised DNA polymerases to replication of structured DNA. DNA Repair (Amst.) 2015, 29, 83–90. [Google Scholar] [CrossRef] [PubMed]
- Gellert, M.; Lipsett, M.N.; Davies, D.R. Helix formation by guanylic acid. Proc. Natl. Acad. Sci. USA 1962, 48, 2013–2018. [Google Scholar] [CrossRef] [PubMed]
- Sen, D.; Gilbert, W. Formation of parallel four-stranded complexes by guanine-rich motifs in DNA and its implications for meiosis. Nature 1988, 334, 364–366. [Google Scholar] [CrossRef] [PubMed]
- Sundquist, W.I.; Klug, A. Telomeric DNA dimerizes by formation of guanine tetrads between hairpin loops. Nature 1989, 342, 825–829. [Google Scholar] [CrossRef] [PubMed]
- Williamson, J.R.; Raghuraman, M.K.; Cech, T.R. Monovalent cation-induced structure of telomeric DNA: The G-quartet model. Cell 1989, 59, 871–880. [Google Scholar] [CrossRef]
- Huppert, J.L.; Balasubramanian, S. Prevalence of quadruplexes in the human genome. Nucleic Acids Res. 2005, 33, 2908–2916. [Google Scholar] [CrossRef] [PubMed]
- Todd, A.K.; Johnston, M.; Neidle, S. Highly prevalent putative quadruplex sequence motifs in human DNA. Nucleic Acids Res. 2005, 33, 2901–2907. [Google Scholar] [CrossRef]
- Schiavone, D.; Guilbaud, G.; Murat, P.; Papadopoulou, C.; Sarkies, P.; Prioleau, M.-N.; Balasubramanian, S.; Sale, J.E. Determinants of G quadruplex-induced epigenetic instability in REV1-deficient cells. The EMBO J. 2014, 33, 2507–2520. [Google Scholar] [CrossRef]
- Wang, L.; Wang, Q.-M.; Wang, Y.-R.; Xi, X.-G.; Hou, X.-M. DNA-unwinding activity of Saccharomyces cerevisiae Pif1 is modulated by thermal stability, folding conformation, and loop lengths of G-quadruplex DNA. J. Biol. Chem. 2018, 293, 18504–18513. [Google Scholar] [CrossRef]
- Chambers, V.S.; Marsico, G.; Boutell, J.M.; Di Antonio, M.; Smith, G.P.; Balasubramanian, S. High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat. Biotechnol. 2015, 33, 877–881. [Google Scholar] [CrossRef] [PubMed]
- Burge, S.; Parkinson, G.N.; Hazel, P.; Todd, A.K.; Neidle, S. Quadruplex DNA: Sequence, topology and structure. Nucleic Acids Res. 2006, 34, 5402–5415. [Google Scholar] [CrossRef] [PubMed]
- Maizels, N.; Gray, L.T. The G4 genome. PLoS Genet. 2013, 9, e1003468. [Google Scholar] [CrossRef] [PubMed]
- Murat, P.; Balasubramanian, S. Existence and consequences of G-quadruplex structures in DNA. Curr. Opin. Genet. Dev. 2014, 25, 22–29. [Google Scholar] [CrossRef] [PubMed]
- Rhodes, D.; Lipps, H.J. G-quadruplexes and their regulatory roles in biology. Nucleic Acids Res. 2015, 43, 8627–8637. [Google Scholar] [CrossRef] [PubMed]
- Tarsounas, M.; Tijsterman, M. Genomes and G-quadruplexes: For better or for worse. J. Mol. Biol. 2013, 425, 4782–4789. [Google Scholar] [CrossRef] [PubMed]
- Siddiqui-Jain, A.; Grand, C.L.; Bearss, D.J.; Hurley, L.H. Direct evidence for a G-quadruplex in a promoter region and its targeting with a small molecule to repress c-MYC transcription. Proc. Natl. Acad. Sci. USA 2002, 99, 11593–11598. [Google Scholar] [CrossRef] [PubMed]
- Paeschke, K.; Simonsson, T.; Postberg, J.; Rhodes, D.; Lipps, H.J. Telomere end-binding proteins control the formation of G-quadruplex DNA structures in vivo. Nat. Struct. Mol. Biol. 2005, 12, 847–854. [Google Scholar] [CrossRef]
- Smith, J.S.; Chen, Q.; Yatsunyk, L.A.; Nicoludis, J.M.; Garcia, M.S.; Kranaster, R.; Balasubramanian, S.; Monchaud, D.; Teulade-Fichou, M.P.; Abramowitz, L.; et al. Rudimentary G-quadruplex-based telomere capping in Saccharomyces cerevisiae. Nat. Struct. Mol. Biol. 2011, 18, 478–485. [Google Scholar] [CrossRef]
- Besnard, E.; Babled, A.; Lapasset, L.; Milhavet, O.; Parrinello, H.; Dantec, C.; Marin, J.M.; Lemaitre, J.M. Unraveling cell type-specific and reprogrammable human replication origin signatures associated with G-quadruplex consensus motifs. Nat. Struct. Mol. Biol. 2012, 19, 837–844. [Google Scholar] [CrossRef]
- Cayrou, C.; Coulombe, P.; Vigneron, A.; Stanojcic, S.; Ganier, O.; Peiffer, I.; Rivals, E.; Puy, A.; Laurent-Chabalier, S.; Desprat, R.; et al. Genome-scale analysis of metazoan replication origins reveals their organization in specific but flexible sites defined by conserved features. Genome Res. 2011, 21, 1438–1449. [Google Scholar] [CrossRef] [PubMed]
- Valton, A.L.; Hassan-Zadeh, V.; Lema, I.; Boggetto, N.; Alberti, P.; Saintomé, C.; Riou, J.F.; Prioleau, M.N. G4 motifs affect origin positioning and efficiency in two vertebrate replicators. EMBO J. 2014, 33, 732–746. [Google Scholar] [CrossRef] [PubMed]
- Woodford, K.J.; Usdin, K.; Weitzmann, M.N. DNA Secondary structures and the evolution of hypervariable tandem arrays. J. Biol. Chem. 1997, 272, 9517–9523. [Google Scholar] [CrossRef]
- Sarkies, P.; Reams, C.; Simpson, L.J.; Sale, J.E. Epigenetic instability due to defective replication of structured DNA. Mol. Cell 2010, 40, 703–713. [Google Scholar] [CrossRef] [PubMed]
- De, S.; Michor, F. DNA secondary structures and epigenetic determinants of cancer genome evolution. Nat. Struct. Mol. Biol. 2011, 18, 950–955. [Google Scholar] [CrossRef] [PubMed]
- Lemmens, B.; van Schendel, R.; Tijsterman, M. Mutagenic consequences of a single G-quadruplex demonstrate mitotic inheritance of DNA replication fork barriers. Nat. Commun. 2015, 6, 8909. [Google Scholar] [CrossRef] [PubMed]
- London, T.B.; Barber, L.J.; Mosedale, G.; Kelly, G.P.; Balasubramanian, S.; Hickson, I.D.; Boulton, S.J.; Hiom, K. FANCJ is a structure-specific DNA helicase associated with the maintenance of genomic G/C tracts. J. Biol. Chem. 2008, 283, 36132–36139. [Google Scholar] [CrossRef]
- Krasilnikova, M.M.; Samadashwily, G.M.; Krasilnikov, A.S.; Mirkin, S.M. Transcription through a simple DNA repeat blocks replication elongation. EMBO J. 1998, 17, 5095–5102. [Google Scholar] [CrossRef]
- Krasilnikova, M.M.; Mirkin, S.M. Replication stalling at Friedreich’s ataxia (GAA)n repeats in vivo. Mol. Cell. Biol. 2004, 24, 2286–2295. [Google Scholar] [CrossRef]
- Chandok, G.S.; Patel, M.P.; Mirkin, S.M.; Krasilnikova, M.M. Effects of Friedreich’s ataxia GAA repeats on DNA replication in mammalian cells. Nucleic Acids Res. 2012, 40, 3964–3974. [Google Scholar] [CrossRef]
- Follonier, C.; Oehler, J.; Herrador, R.; Lopes, M. Friedreich’s ataxia-associated GAA repeats induce replication-fork reversal and unusual molecular junctions. Nat. Struct. Mol. Biol. 2013, 20, 486–494. [Google Scholar] [CrossRef] [PubMed]
- Sfeir, A.; Kosiyatrakul, S.T.; Hockemeyer, D.; MacRae, S.L.; Karlseder, J.; Schildkraut, C.L.; de Lange, T. Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 2009, 138, 90–103. [Google Scholar] [CrossRef] [PubMed]
- Gerhardt, J.; Tomishima, M.J.; Zaninovic, N.; Colak, D.; Yan, Z.; Zhan, Q.; Rosenwaks, Z.; Jaffrey, S.R.; Schildkraut, C.L. The DNA replication program is altered at the FMR1 locus in fragile X embryonic stem cells. Mol. Cell 2014, 53, 19–31. [Google Scholar] [CrossRef] [PubMed]
- Gerhardt, J.; Bhalla, A.D.; Butler, J.S.; Puckett, J.W.; Dervan, P.B.; Rosenwaks, Z.; Napierala, M. Stalled DNA replication forks at the endogenous GAA repeats drive repeat expansion in Friedreich’s ataxia cells. Cell Rep. 2016, 16, 1218–1227. [Google Scholar] [CrossRef] [PubMed]
- Dovrat, D.; Dahan, D.; Sherman, S.; Tsirkas, I.; Elia, N.; Aharoni, A. A live-cell imaging approach for measuring DNA replication rates. Cell Rep. 2018, 24, 252–258. [Google Scholar] [CrossRef]
- Shishkin, A.A.; Voineagu, I.; Matera, R.; Cherng, N.; Chernet, B.T.; Krasilnikova, M.M.; Narayanan, V.; Lobachev, K.S.; Mirkin, S.M. Large-scale expansions of Friedreich’s ataxia GAA repeats in yeast. Mol. Cell 2009, 35, 82–92. [Google Scholar] [CrossRef] [PubMed]
- Kononenko, A.V.; Ebersole, T.; Vasquez, K.M.; Mirkin, S.M. Mechanisms of genetic instability caused by (CGG)n repeats in an experimental mammalian system. Nat. Struct. Mol. Biol. 2018, 25, 669–676. [Google Scholar] [CrossRef]
- Kruisselbrink, E.; Guryev, V.; Brouwer, K.; Pontier, D.B.; Cuppen, E.; Tijsterman, M. Mutagenic capacity of endogenous G4 DNA underlies genome instability in FANCJ-defective C. elegans. Curr. Biol. 2008, 18, 900–905. [Google Scholar] [CrossRef]
- Ribeyre, C.; Lopes, J.; Boulé, J.B.; Piazza, A.; Guédin, A.; Zakian, V.A.; Mergny, J.L.; Nicolas, A. The yeast Pif1 helicase prevents genomic instability caused by G-quadruplex-forming CEB1 sequences in vivo. PLoS Genet. 2009, 5, e1000475. [Google Scholar] [CrossRef]
- Luria, S.E.; Delbrück, M. Mutations of bacteria from virus sensitivity to virus resistance. Genetics 1943, 28, 491–511. [Google Scholar]
- Shastri, N.; Tsai, Y.C.; Hile, S.; Jordan, D.; Powell, B.; Chen, J.; Maloney, D.; Dose, M.; Lo, Y.; Anastassiadis, T.; et al. Genome-wide identification of structure-forming repeats as principal sites of fork collapse upon ATR inhibition. Mol. Cell 2018, 72, 222–238. [Google Scholar] [CrossRef] [PubMed]
- Tubbs, A.; Sridharan, S.; van Wietmarschen, N.; Maman, Y.; Callen, E.; Stanlie, A.; Wu, W.; Wu, X.; Day, A.; Wong, N.; et al. Dual roles of poly(dA:dT) tracts in replication initiation and fork collapse. Cell. 2018, 174, 1127.e19–1142.e19. [Google Scholar] [CrossRef] [PubMed]
- Šviković, S.; Sale, J.E. The effects of replication stress on S phase histone management and epigenetic memory. J. Mol. Biol. 2017, 429, 2011–2029. [Google Scholar] [CrossRef] [PubMed]
- Guilbaud, G.; Murat, P.; Recolin, B.; Campbell, B.C.; Maiter, A.; Sale, J.E.; Balasubramanian, S. Local epigenetic reprogramming induced by G-quadruplex ligands. Nat. Chem. 2017, 9, 1110–1117. [Google Scholar] [CrossRef] [PubMed]
- Wang, G.; Vasquez, K.M. Effects of replication and transcription on DNA structure-related genetic instability. Genes (Basel) 2017, 8, 17. [Google Scholar] [CrossRef] [PubMed]
- Hoede, C.; Denamur, E.; Tenaillon, O. Selection acts on DNA secondary structures to decrease transcriptional mutagenesis. PLoS Genet. 2006, 2, e176. [Google Scholar] [CrossRef] [PubMed]
- Todd, P.A.; Glickman, B.W. Mutational specificity of UV light in Escherichia coli: Indications for a role of DNA secondary structure. Proc. Natl. Acad. Sci. USA 1982, 79, 4123–4127. [Google Scholar] [CrossRef]
- Bacolla, A.; Tainer, J.A.; Vasquez, K.M.; Cooper, D.N. Translocation and deletion breakpoints in cancer genomes are associated with potential non-B DNA-forming sequences. Nucleic Acids Res. 2016, 44, 5673–5688. [Google Scholar] [CrossRef]
- De, S.; Shaknovich, R.; Riester, M.; Elemento, O.; Geng, H.; Kormaksson, M.; Jiang, Y.; Woolcock, B.; Johnson, N.; Polo, J.M.; et al. Aberration in DNA methylation in B-cell lymphomas has a complex origin and increases with disease severity. PLoS Genet. 2013, 9, e1003137. [Google Scholar] [CrossRef]
- Georgakopoulos-Soares, I.; Morganella, S.; Jain, N.; Hemberg, M.; Nik-Zainal, S. Noncanonical secondary structures arising from non-B DNA motifs are determinants of mutagenesis. Genome Res. 2018, 28, 1264–1271. [Google Scholar] [CrossRef]
- Feinberg, A.P.; Ohlsson, R.; Henikoff, S. The epigenetic progenitor origin of human cancer. Nat. Rev. Genet. 2006, 7, 21–33. [Google Scholar] [CrossRef] [PubMed]
- Cheung, I.; Schertzer, M.; Rose, A.; Lansdorp, P.M. Disruption of dog-1 in Caenorhabditis elegans triggers deletions upstream of guanine-rich DNA. Nat. Genet. 2002, 31, 405–409. [Google Scholar] [CrossRef] [PubMed]
- Ding, H.; Schertzer, M.; Wu, X.; Gertsenstein, M.; Selig, S.; Kammori, M.; Pourvali, R.; Poon, S.; Vulto, I.; Chavez, E.; et al. Regulation of murine telomere length by Rtel: An essential gene encoding a helicase-like protein. Cell 2004, 117, 873–886. [Google Scholar] [CrossRef] [PubMed]
- Safa, L.; Gueddouda, N.M.; Thiébaut, F.; Delagoutte, E.; Petruseva, I.; Lavrik, O.; Mendoza, O.; Bourdoncle, A.; Alberti, P.; Riou, J.F.; et al. 5′ to 3′ unfolding directionality of DNA secondary structures by replication protein A: G-quadruplexes and duplexes. J. Biol. Chem. 2016, 291, 21246–21256. [Google Scholar] [CrossRef] [PubMed]
- Dahan, D.; Tsirkas, I.; Dovrat, D.; Sparks, M.A.; Singh, S.P.; Galletto, R.; Aharoni, A. Pif1 is essential for efficient replisome progression through lagging strand G-quadruplex DNA secondary structures. Nucleic Acids Res. 2018, 46, 11847–11857. [Google Scholar] [CrossRef] [PubMed]
- Lopes, J.; Piazza, A.; Bermejo, R.; Kriegsman, B.; Colosio, A.; Teulade-Fichou, M.P.; Foiani, M.; Nicolas, A. G-quadruplex-induced instability during leading-strand replication. EMBO J. 2011, 30, 4033–4046. [Google Scholar] [CrossRef] [PubMed]
- Sarkies, P.; Murat, P.; Phillips, L.G.; Patel, K.J.; Balasubramanian, S.; Sale, J.E. FANCJ coordinates two pathways that maintain epigenetic stability at G-quadruplex DNA. Nucleic Acids Res. 2012, 40, 1485–1498. [Google Scholar] [CrossRef]
- Papadopoulou, C.; Guilbaud, G.; Schiavone, D.; Sale, J.E. Nucleotide pool depletion induces G-quadruplex-dependent perturbation of gene expression. Cell Rep. 2015, 13, 2491–2503. [Google Scholar] [CrossRef]
- Schiavone, D.; Jozwiakowski, S.K.; Romanello, M.; Guilbaud, G.; Guilliam, T.A.; Bailey, L.J.; Sale, J.E.; Doherty, A.J. PrimPol is required for replicative tolerance of G quadruplexes in vertebrate cells. Mol. Cell 2016, 61, 161–169. [Google Scholar] [CrossRef]
- Fenouil, R.; Cauchy, P.; Koch, F.; Descostes, N.; Cabeza, J.Z.; Innocenti, C.; Ferrier, P.; Spicuglia, S.; Gut, M.; Gut, I.; et al. CpG islands and GC content dictate nucleosome depletion in a transcription-independent manner at mammalian promoters. Genome Res. 2012, 22, 2399–2408. [Google Scholar] [CrossRef]
- Wong, H.M.; Huppert, J.L. Stable G-quadruplexes are found outside nucleosome-bound regions. Mol. Biosyst. 2009, 5, 1713–1719. [Google Scholar] [CrossRef] [PubMed]
- Law, M.J.; Lower, K.M.; Voon, H.P.; Hughes, J.R.; Garrick, D.; Viprakasit, V.; Mitson, M.; De Gobbi, M.; Marra, M.; Morris, A.; et al. ATR-X syndrome protein targets tandem repeats and influences allele-specific expression in a size-dependent manner. Cell 2010, 143, 367–378. [Google Scholar] [CrossRef] [PubMed]
- Lewis, P.W.; Elsaesser, S.J.; Noh, K.M.; Stadler, S.C.; Allis, C.D. Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in replication-independent chromatin assembly at telomeres. Proc. Natl. Acad. Sci. USA 2010, 107, 14075–14080. [Google Scholar] [CrossRef] [PubMed]
- Clynes, D.; Jelinska, C.; Xella, B.; Ayyub, H.; Scott, C.; Mitson, M.; Taylor, S.; Higgs, D.R.; Gibbons, R.J. Suppression of the alternative lengthening of telomere pathway by the chromatin remodelling factor ATRX. Nat. Commun. 2015, 6, 7538. [Google Scholar] [CrossRef] [PubMed]
- Mirkin, S.M.; Lyamichev, V.I.; Drushlyak, K.N.; Dobrynin, V.N.; Filippov, S.A.; Frank-Kamenetskii, M.D. DNA H form requires a homopurine-homopyrimidine mirror repeat. Nature 1987, 330, 495–497. [Google Scholar] [CrossRef] [PubMed]
- Sekibo, D.A.T.; Fox, K.R. The effects of DNA supercoiling on G-quadruplex formation. Nucleic Acids Res. 2017, 45, 12069–12079. [Google Scholar] [CrossRef]
- Shirude, P.S.; Okumus, B.; Ying, L.; Ha, T.; Balasubramanian, S. Single-molecule conformational analysis of G-quadruplex formation in the promoter DNA duplex of the proto-oncogene c-kit. J. Am. Chem. Soc. 2007, 129, 7484–7485. [Google Scholar] [CrossRef]
- Bugaut, A.; Balasubramanian, S. A sequence-independent study of the influence of short loop lengths on the stability and topology of intramolecular DNA G-quadruplexes. Biochemistry 2008, 47, 689–697. [Google Scholar] [CrossRef]
- Kireeva, M.L.; Komissarova, N.; Waugh, D.S.; Kashlev, M. The 8-nucleotide-long RNA:DNA hybrid is a primary stability determinant of the RNA polymerase II elongation complex. J. Biol. Chem. 2000, 275, 6530–6536. [Google Scholar] [CrossRef]
- Gómez-González, B.; Aguilera, A. Activation-induced cytidine deaminase action is strongly stimulated by mutations of the THO complex. Proc. Natl. Acad. Sci. USA 2007, 104, 8409–8414. [Google Scholar] [CrossRef]
- Thomas, M.; White, R.L.; Davis, R.W. Hybridization of RNA to double-stranded DNA: Formation of R-loops. Proc. Natl. Acad. Sci. USA 1976, 73, 2294–2298. [Google Scholar] [CrossRef] [PubMed]
- Daniels, G.A.; Lieber, M.R. RNA:DNA complex formation upon transcription of immunoglobulin switch regions: Implications for the mechanism and regulation of class switch recombination. Nucleic Acids Res. 1995, 23, 5006–5011. [Google Scholar] [CrossRef] [PubMed]
- Duquette, M.L.; Handa, P.; Vincent, J.A.; Taylor, A.F.; Maizels, N. Intracellular transcription of G-rich DNAs induces formation of G-loops, novel structures containing G4 DNA. Genes Dev. 2004, 18, 1618–1629. [Google Scholar] [CrossRef] [PubMed]
- Šviković, S.; Crisp, A.; Tan-Wong, S.M.; Guilliam, T.A.; Doherty, A.J.; Proudfoot, N.J.; Guilbaud, G.; Sale, J.E. R-loop formation during S phase is restricted by PrimPol-mediated repriming. EMBO J. 2018. [Google Scholar] [CrossRef]
- Kouzine, F.; Sanford, S.; Elisha-Feil, Z.; Levens, D. The functional response of upstream DNA to dynamic supercoiling in vivo. Nat. Struct. Mol. Biol. 2008, 15, 146–154. [Google Scholar] [CrossRef] [PubMed]
- Jeon, J.H.; Adamcik, J.; Dietler, G.; Metzler, R. Supercoiling induces denaturation bubbles in circular DNA. Phys. Rev. Lett. 2010, 105, 208101. [Google Scholar] [CrossRef]
- Parsa, J.Y.; Ramachandran, S.; Zaheen, A.; Nepal, R.M.; Kapelnikov, A.; Belcheva, A.; Berru, M.; Ronai, D.; Martin, A. Negative supercoiling creates single-stranded patches of DNA that are substrates for AID-mediated mutagenesis. PLoS Genet. 2012, 8, e1002518. [Google Scholar] [CrossRef]
- Romanello, M.; Schiavone, D.; Frey, A.; Sale, J.E. Histone H3.3 promotes IgV gene diversification by enhancing formation of AID-accessible single-stranded DNA. EMBO J. 2016, 35, 1452–1464. [Google Scholar] [CrossRef]
- Dayn, A.; Samadashwily, G.M.; Mirkin, S.M. Intramolecular DNA triplexes: Unusual sequence requirements and influence on DNA polymerization. Proc. Natl. Acad. Sci. USA 1992, 89, 11406–11410. [Google Scholar] [CrossRef]
- Lilley, D.M. The inverted repeat as a recognizable structural feature in supercoiled DNA molecules. Proc. Natl. Acad. Sci. USA 1980, 77, 6468–6472. [Google Scholar] [CrossRef]
- Panayotatos, N.; Wells, R.D. Cruciform structures in supercoiled DNA. Nature 1981, 289, 466–470. [Google Scholar] [CrossRef] [PubMed]
- Grabczyk, E.; Fishman, M.C. A long purine-pyrimidine homopolymer acts as a transcriptional diode. J. Biol. Chem. 1995, 270, 1791–1797. [Google Scholar] [CrossRef] [PubMed]
- Grabczyk, E.; Usdin, K. The GAA*TTC triplet repeat expanded in Friedreich’s ataxia impedes transcription elongation by T7 RNA polymerase in a length and supercoil dependent manner. Nucleic Acids Res. 2000, 28, 2815–2822. [Google Scholar] [CrossRef] [PubMed]
- Kouzine, F.; Liu, J.; Sanford, S.; Chung, H.J.; Levens, D. The dynamic response of upstream DNA to transcription-generated torsional stress. Nat. Struct. Mol. Biol. 2004, 11, 1092–1100. [Google Scholar] [CrossRef] [PubMed]
- Yadav, P.; Harcy, V.; Argueso, J.L.; Dominska, M.; Jinks-Robertson, S.; Kim, N. Topoisomerase I plays a critical role in suppressing genome instability at a highly transcribed G-quadruplex-forming sequence. PLoS Genet. 2014, 10, e1004839. [Google Scholar] [CrossRef] [PubMed]
- Biffi, G.; Tannahill, D.; McCafferty, J.; Balasubramanian, S. Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 2013, 5, 182–186. [Google Scholar] [CrossRef] [PubMed]
- Moyer, S.E.; Lewis, P.W.; Botchan, M.R. Isolation of the Cdc45/Mcm2-7/GINS (CMG) complex, a candidate for the eukaryotic DNA replication fork helicase. Proc. Natl. Acad. Sci. USA 2006, 103, 10236–10241. [Google Scholar] [CrossRef] [PubMed]
- Georgescu, R.; Yuan, Z.; Bai, L.; de Luna Almeida Santos, R.; Sun, J.; Zhang, D.; Yurieva, O.; Li, H.; O’Donnell, M.E. Structure of eukaryotic CMG helicase at a replication fork and implications to replisome architecture and origin initiation. Proc. Natl. Acad. Sci. USA 2017, 114, E697–E706. [Google Scholar] [CrossRef] [PubMed]
- Noguchi, Y.; Yuan, Z.; Bai, L.; Schneider, S.; Zhao, G.; Stillman, B.; Speck, C.; Li, H. Cryo-EM structure of Mcm2-7 double hexamer on DNA suggests a lagging-strand DNA extrusion model. Proc. Natl. Acad. Sci. USA 2017, 114, E9529–E9538. [Google Scholar] [CrossRef]
- Douglas, M.E.; Ali, F.A.; Costa, A.; Diffley, J.F.X. The mechanism of eukaryotic CMG helicase activation. Nature 2018, 555, 265–268. [Google Scholar] [CrossRef] [PubMed]
- Yuan, Z.; Bai, L.; Sun, J.; Georgescu, R.; Liu, J.; O’Donnell, M.E.; Li, H. Structure of the eukaryotic replicative CMG helicase suggests a pumpjack motion for translocation. Nat. Struct. Mol. Biol. 2016, 23, 217–224. [Google Scholar] [CrossRef] [PubMed]
- Fu, Y.V.; Yardimci, H.; Long, D.T.; Ho, T.V.; Guainazzi, A.; Bermudez, V.P.; Hurwitz, J.; van Oijen, A.; Schärer, O.D.; Walter, J.C. Selective bypass of a lagging strand roadblock by the eukaryotic replicative DNA helicase. Cell 2011, 146, 931–941. [Google Scholar] [CrossRef] [PubMed]
- Langston, L.D.; Mayle, R.; Schauer, G.D.; Yurieva, O.; Zhang, D.; Yao, N.Y.; Georgescu, R.E.; O’Donnell, M.E. Mcm10 promotes rapid isomerization of CMG-DNA for replisome bypass of lagging strand DNA blocks. eLife 2017, 6, e29118. [Google Scholar] [CrossRef] [PubMed]
- Salas, T.R.; Petruseva, I.; Lavrik, O.; Bourdoncle, A.; Mergny, J.L.; Favre, A.; Saintomé, C. Human replication protein A unfolds telomeric G-quadruplexes. Nucleic Acids Res. 2006, 34, 4857–4865. [Google Scholar] [CrossRef] [PubMed]
- Remus, D.; Beuron, F.; Tolun, G.; Griffith, J.D.; Morris, E.P.; Diffley, J.F. Concerted loading of Mcm2-7 double hexamers around DNA during DNA replication origin licensing. Cell 2009, 139, 719–730. [Google Scholar] [CrossRef] [PubMed]
- Evrin, C.; Clarke, P.; Zech, J.; Lurz, R.; Sun, J.; Uhle, S.; Li, H.; Stillman, B.; Speck, C. A double-hexameric MCM2-7 complex is loaded onto origin DNA during licensing of eukaryotic DNA replication. Proc. Natl. Acad. Sci. USA 2009, 106, 20240–20245. [Google Scholar] [CrossRef] [PubMed]
- Langston, L.; O’Donnell, M. Action of CMG with strand-specific DNA blocks supports an internal unwinding mode for the eukaryotic replicative helicase. eLife 2017, 6, e23449. [Google Scholar] [CrossRef] [PubMed]
- Amrane, S.; Adrian, M.; Heddi, B.; Serero, A.; Nicolas, A.; Mergny, J.L.; Phan, A.T. Formation of pearl-necklace monomorphic G-quadruplexes in the human CEB25 minisatellite. J. Am. Chem. Soc. 2012, 134, 5807–5816. [Google Scholar] [CrossRef] [PubMed]
- Do, N.Q.; Lim, K.W.; Teo, M.H.; Heddi, B.; Phan, A.T. Stacking of G-quadruplexes: NMR structure of a G-rich oligonucleotide with potential anti-HIV and anticancer activity. Nucleic Acids Res. 2011, 39, 9448–9457. [Google Scholar] [CrossRef]
- Heddi, B.; Phan, A.T. Structure of human telomeric DNA in crowded solution. J. Am. Chem. Soc. 2011, 133, 9824–9833. [Google Scholar] [CrossRef]
- Wang, Y.; Patel, D.J. Solution structure of a parallel-stranded G-quadruplex DNA. J. Mol. Biol. 1993, 234, 1171–1183. [Google Scholar] [CrossRef] [PubMed]
- Yardimci, H.; Wang, X.; Loveland, A.B.; Tappin, I.; Rudner, D.Z.; Hurwitz, J.; van Oijen, A.M.; Walter, J.C. Bypass of a protein barrier by a replicative DNA helicase. Nature 2012, 492, 205–209. [Google Scholar] [CrossRef] [PubMed]
- Sparks, J.L.; Chistol, G.; Gao, A.O.; Räschle, M.; Larsen, N.B.; Mann, M.; Duxin, J.P.; Walter, J.C. The CMG helicase bypasses DNA-protein cross-links to facilitate their repair. Cell 2019, 176, 167–181.e21. [Google Scholar] [CrossRef] [PubMed]
- O’Donnell, M.E.; Li, H. The ring-shaped hexameric helicases that function at DNA replication forks. Nat. Struct. Mol. Biol. 2018, 25, 122–130. [Google Scholar] [CrossRef] [PubMed]
- Zou, L.; Elledge, S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003, 300, 1542–1548. [Google Scholar] [CrossRef] [PubMed]
- Byun, T.S.; Pacek, M.; Yee, M.C.; Walter, J.C.; Cimprich, K.A. Functional uncoupling of MCM helicase and DNA polymerase activities activates the ATR-dependent checkpoint. Genes Dev. 2005, 19, 1040–1052. [Google Scholar] [CrossRef] [PubMed]
- Audry, J.; Maestroni, L.; Delagoutte, E.; Gauthier, T.; Nakamura, T.M.; Gachet, Y.; Saintomé, C.; Géli, V.; Coulon, S. RPA prevents G-rich structure formation at lagging-strand telomeres to allow maintenance of chromosome ends. EMBO J. 2015, 34, 1942–1958. [Google Scholar] [CrossRef]
- Eddy, S.; Ketkar, A.; Zafar, M.K.; Maddukuri, L.; Choi, J.Y.; Eoff, R.L. Human Rev1 polymerase disrupts G-quadruplex DNA. Nucleic Acids Res. 2014, 42, 3272–3285. [Google Scholar] [CrossRef]
- Budhathoki, J.B.; Ray, S.; Urban, V.; Janscak, P.; Yodh, J.G.; Balci, H. RecQ-core of BLM unfolds telomeric G-quadruplex in the absence of ATP. Nucleic Acids Res. 2014, 42, 11528–11545. [Google Scholar] [CrossRef]
- Chen, M.C.; Tippana, R.; Demeshkina, N.A.; Murat, P.; Balasubramanian, S.; Myong, S.; Ferré-D’Amaré, A.R. Structural basis of G-quadruplex unfolding by the DEAH/RHA helicase DHX36. Nature 2018, 558, 465–469. [Google Scholar] [CrossRef]
- Chen, W.F.; Rety, S.; Guo, H.L.; Dai, Y.X.; Wu, W.Q.; Liu, N.N.; Auguin, D.; Liu, Q.W.; Hou, X.M.; Dou, S.X.; et al. Molecular mechanistic insights into Drosophila DHX36-Mediated G-quadruplex unfolding: A structure-based model. Structure 2018, 26, 403–415.e4. [Google Scholar] [CrossRef] [PubMed]
- Yangyuoru, P.M.; Bradburn, D.A.; Liu, Z.; Xiao, T.S.; Russell, R. The G-quadruplex (G4) resolvase DHX36 efficiently and specifically disrupts DNA G4s via a translocation-based helicase mechanism. J. Biol. Chem. 2018, 293, 1924–1932. [Google Scholar] [CrossRef] [PubMed]
- Brosh, R.M. DNA helicases involved in DNA repair and their roles in cancer. Nat. Rev. Cancer 2013, 13, 542–558. [Google Scholar] [CrossRef] [PubMed]
- Maizels, N. G4-associated human diseases. EMBO Rep. 2015, 16, 910–922. [Google Scholar] [CrossRef] [PubMed]
- Singleton, M.R.; Dillingham, M.S.; Wigley, D.B. Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 2007, 76, 23–50. [Google Scholar] [CrossRef]
- Bae, S.H.; Choi, E.; Lee, K.H.; Park, J.S.; Lee, S.H.; Seo, Y.S. Dna2 of Saccharomyces cerevisiae possesses a single-stranded DNA-specific endonuclease activity that is able to act on double-stranded DNA in the presence of ATP. J. Biol. Chem. 1998, 273, 26880–26890. [Google Scholar] [CrossRef] [PubMed]
- Budd, M.E.; Campbell, J.L. A yeast gene required for DNA replication encodes a protein with homology to DNA helicases. Proc. Natl. Acad. Sci. USA 1995, 92, 7642–7646. [Google Scholar] [CrossRef] [PubMed]
- Masuda-Sasa, T.; Polaczek, P.; Peng, X.P.; Chen, L.; Campbell, J.L. Processing of G4 DNA by Dna2 helicase/nuclease and replication protein A (RPA) provides insights into the mechanism of Dna2/RPA substrate recognition. J. Biol. Chem. 2008, 283, 24359–24373. [Google Scholar] [CrossRef]
- Zhou, C.; Pourmal, S.; Pavletich, N.P. Dna2 nuclease-helicase structure, mechanism and regulation by Rpa. eLife 2015, 4, e09832. [Google Scholar] [CrossRef]
- Lin, W.; Sampathi, S.; Dai, H.; Liu, C.; Zhou, M.; Hu, J.; Huang, Q.; Campbell, J.; Shin-Ya, K.; Zheng, L.; et al. Mammalian DNA2 helicase/nuclease cleaves G-quadruplex DNA and is required for telomere integrity. EMBO J. 2013, 32, 1425–1439. [Google Scholar] [CrossRef] [PubMed]
- Bochman, M.L.; Sabouri, N.; Zakian, V.A. Unwinding the functions of the Pif1 family helicases. DNA Repair (Amst.) 2010, 9, 237–249. [Google Scholar] [CrossRef] [PubMed]
- Liu, N.N.; Duan, X.L.; Ai, X.; Yang, Y.T.; Li, M.; Dou, S.X.; Rety, S.; Deprez, E.; Xi, X.G. The Bacteroides sp. 3_1_23 Pif1 protein is a multifunctional helicase. Nucleic Acids Res. 2015, 43, 8942–8954. [Google Scholar] [CrossRef] [PubMed]
- Mendoza, O.; Gueddouda, N.M.; Boulé, J.B.; Bourdoncle, A.; Mergny, J.L. A fluorescence-based helicase assay: Application to the screening of G-quadruplex ligands. Nucleic Acids Res. 2015, 43, e71. [Google Scholar] [CrossRef]
- Paeschke, K.; Bochman, M.L.; Garcia, P.D.; Cejka, P.; Friedman, K.L.; Kowalczykowski, S.C.; Zakian, V.A. Pif1 family helicases suppress genome instability at G-quadruplex motifs. Nature 2013, 497, 458–462. [Google Scholar] [CrossRef] [PubMed]
- Sanders, C.M. Human Pif1 helicase is a G-quadruplex DNA-binding protein with G-quadruplex DNA-unwinding activity. Biochem. J. 2010, 430, 119–128. [Google Scholar] [CrossRef]
- Wallgren, M.; Mohammad, J.B.; Yan, K.P.; Pourbozorgi-Langroudi, P.; Ebrahimi, M.; Sabouri, N. G-rich telomeric and ribosomal DNA sequences from the fission yeast genome form stable G-quadruplex DNA structures in vitro and are unwound by the Pfh1 DNA helicase. Nucleic Acids Res. 2016, 44, 6213–6231. [Google Scholar] [CrossRef]
- Zhou, R.; Zhang, J.; Bochman, M.L.; Zakian, V.A.; Ha, T. Periodic DNA patrolling underlies diverse functions of Pif1 on R-loops and G-rich DNA. eLife 2014, 3, e02190. [Google Scholar] [CrossRef] [PubMed]
- Hou, X.-M.; Wu, W.-Q.; Duan, X.-L.; Liu, N.-N.; Li, H.-H.; Fu, J.; Dou, S.-X.; Li, M.; Xi, X.-G. Molecular mechanism of G-quadruplex unwinding helicase: Sequential and repetitive unfolding of G-quadruplex by Pif1 helicase. Biochem. J. 2015, 466, 189–199. [Google Scholar] [CrossRef] [PubMed]
- Li, J.H.; Lin, W.X.; Zhang, B.; Nong, D.G.; Ju, H.P.; Ma, J.B.; Xu, C.H.; Ye, F.F.; Xi, X.G.; Li, M.; et al. Pif1 is a force-regulated helicase. Nucleic Acids Res. 2016, 44, 4330–4339. [Google Scholar] [CrossRef] [PubMed]
- Byrd, A.K.; Bell, M.R.; Raney, K.D. Pif1 helicase unfolding of G-quadruplex DNA is highly dependent on sequence and reaction conditions. J. Biol. Chem. 2018, 293, 17792–17802. [Google Scholar] [CrossRef] [PubMed]
- Boulé, J.B.; Zakian, V.A. Roles of Pif1-like helicases in the maintenance of genomic stability. Nucleic Acids Res. 2006, 34, 4147–4153. [Google Scholar] [CrossRef] [PubMed]
- Futami, K.; Shimamoto, A.; Furuichi, Y. Mitochondrial and nuclear localization of human Pif1 helicase. Biol. Pharm. Bull. 2007, 30, 1685–1692. [Google Scholar] [CrossRef] [PubMed]
- McDonald, K.R.; Sabouri, N.; Webb, C.J.; Zakian, V.A. The Pif1 family helicase Pfh1 facilitates telomere replication and has an RPA-dependent role during telomere lengthening. DNA Repair (Amst.) 2014, 24, 80–86. [Google Scholar] [CrossRef]
- Bannwarth, S.; Berg-Alonso, L.; Augé, G.; Fragaki, K.; Kolesar, J.E.; Lespinasse, F.; Lacas-Gervais, S.; Burel-Vandenbos, F.; Villa, E.; Belmonte, F.; et al. Inactivation of Pif1 helicase causes a mitochondrial myopathy in mice. Mitochondrion 2016, 30, 126–137. [Google Scholar] [CrossRef]
- McDonald, K.R.; Guise, A.J.; Pourbozorgi-Langroudi, P.; Cristea, I.M.; Zakian, V.A.; Capra, J.A.; Sabouri, N. Pfh1 is an accessory replicative helicase that interacts with the replisome to facilitate fork progression and preserve genome integrity. PLoS Genet. 2016, 12, e1006238. [Google Scholar] [CrossRef] [PubMed]
- Ellis, N.A.; Groden, J.; Ye, T.Z.; Straughen, J.; Lennon, D.J.; Ciocci, S.; Proytcheva, M.; German, J. The Bloom’s syndrome gene product is homologous to RecQ helicases. Cell 1995, 83, 655–666. [Google Scholar] [CrossRef]
- Yu, C.-E.; Oshima, J.; Fu, Y.-H.; Wijsman, E.M.; Hisama, F.; Alisch, R.; Matthews, S.; Nakura, J.; Miki, T.; Ouais, S.; et al. Positional cloning of the Werner’s syndrome gene. Science 1996, 272, 258–262. [Google Scholar] [CrossRef] [PubMed]
- Huber, M.D.; Duquette, M.L.; Shiels, J.C.; Maizels, N. A conserved G4 DNA binding domain in RecQ family helicases. J. Mol. Biol. 2006, 358, 1071–1080. [Google Scholar] [CrossRef] [PubMed]
- Voter, A.F.; Qiu, Y.; Tippana, R.; Myong, S.; Keck, J.L. A guanine-flipping and sequestration mechanism for G-quadruplex unwinding by RecQ helicases. Nat. Commun. 2018, 9, 4201. [Google Scholar] [CrossRef]
- Sun, H.; Karow, J.K.; Hickson, I.D.; Maizels, N. The Bloom’s syndrome helicase unwinds G4 DNA. J. Biol. Chem. 1998, 273, 27587–27592. [Google Scholar] [CrossRef]
- Huber, M.D.; Lee, D.C.; Maizels, N. G4 DNA unwinding by BLM and Sgs1p: Substrate specificity and substrate-specific inhibition. Nucleic Acids Res. 2002, 30, 3954–3961. [Google Scholar] [CrossRef]
- Mohaghegh, P.; Karow, J.K.; Brosh, R.M.; Bohr, V.A.; Hickson, I.D. The Bloom’s and Werner’s syndrome proteins are DNA structure-specific helicases. Nucleic Acids Res. 2001, 29, 2843–2849. [Google Scholar] [CrossRef] [PubMed]
- Wang, Q.; Liu, J.Q.; Chen, Z.; Zheng, K.W.; Chen, C.Y.; Hao, Y.H.; Tan, Z. G-quadruplex formation at the 3′ end of telomere DNA inhibits its extension by telomerase, polymerase and unwinding by helicase. Nucleic Acids Res. 2011, 39, 6229–6237. [Google Scholar] [CrossRef] [PubMed]
- Liu, J.Q.; Chen, C.Y.; Xue, Y.; Hao, Y.H.; Tan, Z. G-quadruplex hinders translocation of BLM helicase on DNA: A real-time fluorescence spectroscopic unwinding study and comparison with duplex substrates. J. Am. Chem. Soc. 2010, 132, 10521–10527. [Google Scholar] [CrossRef] [PubMed]
- Li, J.-L.; Harrison, R.J.; Reszka, A.P.; Brosh, R.M.; Bohr, V.A.; Neidle, S.; Hickson, I.D. Inhibition of the Bloom’s and Werner’s syndrome helicases by G-quadruplex interacting ligands. Biochemistry 2001, 40, 15194–15202. [Google Scholar] [CrossRef] [PubMed]
- Wu, W.Q.; Hou, X.M.; Li, M.; Dou, S.X.; Xi, X.G. BLM unfolds G-quadruplexes in different structural environments through different mechanisms. Nucleic Acids Res. 2015, 43, 4614–4626. [Google Scholar] [CrossRef] [PubMed]
- Budhathoki, J.B.; Stafford, E.J.; Yodh, J.G.; Balci, H. ATP-dependent G-quadruplex unfolding by Bloom helicase exhibits low processivity. Nucleic Acids Res. 2015, 43, 5961–5970. [Google Scholar] [CrossRef]
- Brosh, R.M.; Li, J.L.; Kenny, M.K.; Karow, J.K.; Cooper, M.P.; Kureekattil, R.P.; Hickson, I.D.; Bohr, V.A. Replication protein A physically interacts with the Bloom’s syndrome protein and stimulates its helicase activity. J. Biol. Chem. 2000, 275, 23500–23508. [Google Scholar] [CrossRef]
- Wu, W.; Rokutanda, N.; Takeuchi, J.; Lai, Y.; Maruyama, R.; Togashi, Y.; Nishikawa, H.; Arai, N.; Miyoshi, Y.; Suzuki, N.; et al. HERC2 Facilitates BLM and WRN Helicase Complex Interaction with RPA to Suppress G-Quadruplex DNA. Cancer Res. 2018, 78, 6371–6385. [Google Scholar] [CrossRef]
- Youds, J.L.; O’Neil, N.J.; Rose, A.M. Homologous recombination is required for genome stability in the absence of DOG-1 in Caenorhabditis elegans. Genetics 2006, 173, 697–708. [Google Scholar] [CrossRef]
- Tang, W.; Robles, A.I.; Beyer, R.P.; Gray, L.T.; Nguyen, G.H.; Oshima, J.; Maizels, N.; Harris, C.C.; Monnat, R.J. The Werner syndrome RECQ helicase targets G4 DNA in human cells to modulate transcription. Hum. Mol. Genet. 2016, 25, 2060–2069. [Google Scholar] [CrossRef]
- Nguyen, G.H.; Tang, W.; Robles, A.I.; Beyer, R.P.; Gray, L.T.; Welsh, J.A.; Schetter, A.J.; Kumamoto, K.; Wang, X.W.; Hickson, I.D.; et al. Regulation of gene expression by the BLM helicase correlates with the presence of G-quadruplex DNA motifs. Proc. Natl. Acad. Sci. USA 2014, 111, 9905–9910. [Google Scholar] [CrossRef] [PubMed]
- Johnson, J.E.; Cao, K.; Ryvkin, P.; Wang, L.S.; Johnson, F.B. Altered gene expression in the Werner and Bloom syndromes is associated with sequences having G-quadruplex forming potential. Nucleic Acids Res. 2010, 38, 1114–1122. [Google Scholar] [CrossRef] [PubMed]
- Hershman, S.G.; Chen, Q.; Lee, J.Y.; Kozak, M.L.; Yue, P.; Wang, L.S.; Johnson, F.B. Genomic distribution and functional analyses of potential G-quadruplex-forming sequences in Saccharomyces cerevisiae. Nucleic Acids Res. 2008, 36, 144–156. [Google Scholar] [CrossRef] [PubMed]
- Opresko, P.L.; Mason, P.A.; Podell, E.R.; Lei, M.; Hickson, I.D.; Cech, T.R.; Bohr, V.A. POT1 stimulates RecQ helicases WRN and BLM to unwind telomeric DNA substrates. J. Biol. Chem. 2005, 280, 32069–32080. [Google Scholar] [CrossRef]
- Vannier, J.B.; Pavicic-Kaltenbrunner, V.; Petalcorin, M.I.; Ding, H.; Boulton, S.J. RTEL1 dismantles T loops and counteracts telomeric G4-DNA to maintain telomere integrity. Cell 2012, 149, 795–806. [Google Scholar] [CrossRef] [PubMed]
- Muftuoglu, M.; Kulikowicz, T.; Beck, G.; Lee, J.W.; Piotrowski, J.; Bohr, V.A. Intrinsic ssDNA annealing activity in the C-terminal region of WRN. Biochemistry 2008, 47, 10247–10254. [Google Scholar] [CrossRef]
- Shen, J.C.; Gray, M.D.; Oshima, J.; Kamath-Loeb, A.S.; Fry, M.; Loeb, L.A. Werner syndrome protein. I. DNA helicase and DNA exonuclease reside on the same polypeptide. J. Biol. Chem. 1998, 273, 34139–34144. [Google Scholar] [CrossRef]
- Fry, M.; Loeb, L.A. Human werner syndrome DNA helicase unwinds tetrahelical structures of the fragile X syndrome repeat sequence d(CGG)n. J. Biol. Chem. 1999, 274, 12797–12802. [Google Scholar] [CrossRef]
- Brosh, R.M.; Waheed, J.; Sommers, J.A. Biochemical characterization of the DNA substrate specificity of Werner syndrome helicase. J. Biol. Chem. 2002, 277, 23236–23245. [Google Scholar] [CrossRef]
- Lee, M.; Shin, S.; Uhm, H.; Hong, H.; Kirk, J.; Hyun, K.; Kulikowicz, T.; Kim, J.; Ahn, B.; Bohr, V.A.; et al. Multiple RPAs make WRN syndrome protein a superhelicase. Nucleic Acids Res. 2018, 46, 4689–4698. [Google Scholar] [CrossRef] [PubMed]
- Brosh, R.M.; Orren, D.K.; Nehlin, J.O.; Ravn, P.H.; Kenny, M.K.; Machwe, A.; Bohr, V.A. Functional and physical interaction between WRN helicase and human replication protein A. J. Biol. Chem. 1999, 274, 18341–18350. [Google Scholar] [CrossRef] [PubMed]
- Von Kobbe, C.; Karmakar, P.; Dawut, L.; Opresko, P.; Zeng, X.; Brosh, R.M.; Hickson, I.D.; Bohr, V.A. Colocalization, physical, and functional interaction between Werner and Bloom syndrome proteins. J. Biol. Chem. 2002, 277, 22035–22044. [Google Scholar] [CrossRef] [PubMed]
- Kamath-Loeb, A.S.; Loeb, L.A.; Johansson, E.; Burgers, P.M.; Fry, M. Interactions between the Werner syndrome helicase and DNA polymerase δ specifically facilitate copying of tetraplex and hairpin structures of the d(CGG)n trinucleotide repeat sequence. J. Biol. Chem. 2001, 276, 16439–16446. [Google Scholar] [CrossRef] [PubMed]
- Damerla, R.R.; Knickelbein, K.E.; Strutt, S.; Liu, F.J.; Wang, H.; Opresko, P.L. Werner syndrome protein suppresses the formation of large deletions during the replication of human telomeric sequences. Cell Cycle 2012, 11, 3036–3044. [Google Scholar] [CrossRef] [PubMed]
- Crabbe, L.; Verdun, R.E.; Haggblom, C.I.; Karlseder, J. Defective telomere lagging strand synthesis in cells lacking WRN helicase activity. Science 2004, 306, 1951–1953. [Google Scholar] [CrossRef] [PubMed]
- Opresko, P.L.; Otterlei, M.; Graakjaer, J.; Bruheim, P.; Dawut, L.; Kølvraa, S.; May, A.; Seidman, M.M.; Bohr, V.A. The Werner syndrome helicase and exonuclease cooperate to resolve telomeric D loops in a manner regulated by TRF1 and TRF2. Mol. Cell 2004, 14, 763–774. [Google Scholar] [CrossRef] [PubMed]
- Chang, S.; Multani, A.S.; Cabrera, N.G.; Naylor, M.L.; Laud, P.; Lombard, D.; Pathak, S.; Guarente, L.; DePinho, R.A. Essential role of limiting telomeres in the pathogenesis of Werner syndrome. Nat. Genet. 2004, 36, 877–882. [Google Scholar] [CrossRef] [PubMed]
- Du, X.; Shen, J.; Kugan, N.; Furth, E.E.; Lombard, D.B.; Cheung, C.; Pak, S.; Luo, G.; Pignolo, R.J.; DePinho, R.A.; et al. Telomere shortening exposes functions for the mouse Werner and Bloom syndrome genes. Mol. Cell. Biol. 2004, 24, 8437–8446. [Google Scholar] [CrossRef]
- Laud, P.R.; Multani, A.S.; Bailey, S.M.; Wu, L.; Ma, J.; Kingsley, C.; Lebel, M.; Pathak, S.; DePinho, R.A.; Chang, S. Elevated telomere-telomere recombination in WRN-deficient, telomere dysfunctional cells promotes escape from senescence and engagement of the ALT pathway. Genes Dev. 2005, 19, 2560–2570. [Google Scholar] [CrossRef]
- White, M.F.; Dillingham, M.S. Iron-sulphur clusters in nucleic acid processing enzymes. Curr. Opin. Struct. Biol. 2012, 22, 94–100. [Google Scholar] [CrossRef] [PubMed]
- Wu, Y.; Brosh, R.M. DNA helicase and helicase-nuclease enzymes with a conserved iron-sulfur cluster. Nucleic Acids Res. 2012, 40, 4247–4260. [Google Scholar] [CrossRef] [PubMed]
- Levitus, M.; Waisfisz, Q.; Godthelp, B.C.; de Vries, Y.; Hussain, S.; Wiegant, W.W.; Elghalbzouri-Maghrani, E.; Steltenpool, J.; Rooimans, M.A.; Pals, G.; et al. The DNA helicase BRIP1 is defective in Fanconi anemia complementation group J. Nat. Genet. 2005, 37, 934–935. [Google Scholar] [CrossRef] [PubMed]
- Litman, R.; Peng, M.; Jin, Z.; Zhang, F.; Zhang, J.; Powell, S.; Andreassen, P.R.; Cantor, S.B. BACH1 is critical for homologous recombination and appears to be the Fanconi anemia gene product FANCJ. Cancer Cell 2005, 8, 255–265. [Google Scholar] [CrossRef] [PubMed]
- Gupta, R.; Sharma, S.; Sommers, J.A.; Kenny, M.K.; Cantor, S.B.; Brosh, R.M. FANCJ (BACH1) helicase forms DNA damage inducible foci with replication protein A and interacts physically and functionally with the single-stranded DNA-binding protein. Blood 2007, 110, 2390–2398. [Google Scholar] [CrossRef] [PubMed]
- Wu, Y.; Shin-Ya, K.; Brosh, R.M. FANCJ helicase defective in Fanconi anemia and breast cancer unwinds G-quadruplex DNA to defend genomic stability. Mol. Cell. Biol. 2008, 28, 4116–4128. [Google Scholar] [CrossRef] [PubMed]
- Wu, C.G.; Spies, M. G-quadruplex recognition and remodeling by the FANCJ helicase. Nucleic Acids Res. 2016, 44, 8742–8753. [Google Scholar] [CrossRef] [PubMed]
- Alabert, C.; Bukowski-Wills, J.C.; Lee, S.B.; Kustatscher, G.; Nakamura, K.; de Lima Alves, F.; Menard, P.; Mejlvang, J.; Rappsilber, J.; Groth, A. Nascent chromatin capture proteomics determines chromatin dynamics during DNA replication and identifies unknown fork components. Nat. Cell Biol. 2014, 16, 281–293. [Google Scholar] [CrossRef] [PubMed]
- Sirbu, B.M.; McDonald, W.H.; Dungrawala, H.; Badu-Nkansah, A.; Kavanaugh, G.M.; Chen, Y.; Tabb, D.L.; Cortez, D. Identification of proteins at active, stalled, and collapsed replication forks using isolation of proteins on nascent DNA (iPOND) coupled with mass spectrometry. J. Biol. Chem. 2013, 288, 31458–31467. [Google Scholar] [CrossRef]
- Déjardin, J.; Kingston, R.E. Purification of proteins associated with specific genomic Loci. Cell 2009, 136, 175–186. [Google Scholar] [CrossRef]
- Schwab, R.A.; Nieminuszczy, J.; Shin-ya, K.; Niedzwiedz, W. FANCJ couples replication past natural fork barriers with maintenance of chromatin structure. J. Cell Biol. 2013, 201, 33–48. [Google Scholar] [CrossRef] [PubMed]
- Castillo Bosch, P.; Segura-Bayona, S.; Koole, W.; van Heteren, J.T.; Dewar, J.M.; Tijsterman, M.; Knipscheer, P. FANCJ promotes DNA synthesis through G-quadruplex structures. EMBO J. 2014, 33, 2521–2533. [Google Scholar] [CrossRef] [PubMed]
- Youds, J.L.; Barber, L.J.; Ward, J.D.; Collis, S.J.; O’Neil, N.J.; Boulton, S.J.; Rose, A.M. DOG-1 is the Caenorhabditis elegans BRIP1/FANCJ homologue and functions in interstrand cross-link repair. Mol. Cell. Biol. 2008, 28, 1470–1479. [Google Scholar] [CrossRef] [PubMed]
- Inoue, A.; Li, T.; Roby, S.K.; Valentine, M.B.; Inoue, M.; Boyd, K.; Kidd, V.J.; Lahti, J.M. Loss of ChlR1 helicase in mouse causes lethality due to the accumulation of aneuploid cells generated by cohesion defects and placental malformation. Cell Cycle 2007, 6, 1646–1654. [Google Scholar] [CrossRef] [PubMed]
- Alkhunaizi, E.; Shaheen, R.; Bharti, S.K.; Joseph-George, A.M.; Chong, K.; Abdel-Salam, G.M.H.; Alowain, M.; Blaser, S.I.; Papsin, B.C.; Butt, M.; et al. Warsaw breakage syndrome: Further clinical and genetic delineation. Am. J. Med. Genet. A 2018, 176, 2404–2418. [Google Scholar] [CrossRef] [PubMed]
- Bharti, S.K.; Sommers, J.A.; George, F.; Kuper, J.; Hamon, F.; Shin-ya, K.; Teulade-Fichou, M.P.; Kisker, C.; Brosh, R.M. Specialization among iron-sulfur cluster helicases to resolve G-quadruplex DNA structures that threaten genomic stability. J. Biol. Chem. 2013, 288, 28217–28229. [Google Scholar] [CrossRef] [PubMed]
- Wu, Y.; Sommers, J.A.; Khan, I.; de Winter, J.P.; Brosh, R.M. Biochemical characterization of Warsaw breakage syndrome helicase. J. Biol. Chem. 2012, 287, 1007–1021. [Google Scholar] [CrossRef] [PubMed]
- Guo, M.; Hundseth, K.; Ding, H.; Vidhyasagar, V.; Inoue, A.; Nguyen, C.H.; Zain, R.; Lee, J.S.; Wu, Y. A distinct triplex DNA unwinding activity of ChlR1 helicase. J. Biol. Chem. 2015, 290, 5174–5189. [Google Scholar] [CrossRef]
- Mayer, M.L.; Pot, I.; Chang, M.; Xu, H.; Aneliunas, V.; Kwok, T.; Newitt, R.; Aebersold, R.; Boone, C.; Brown, G.W.; et al. Identification of protein complexes required for efficient sister chromatid cohesion. Mol. Biol. Cell 2004, 15, 1736–1745. [Google Scholar] [CrossRef]
- Skibbens, R.V. Chl1p, a DNA helicase-like protein in budding yeast, functions in sister-chromatid cohesion. Genetics 2004, 166, 33–42. [Google Scholar] [CrossRef]
- Gerring, S.L.; Spencer, F.; Hieter, P. The CHL 1 (CTF 1) gene product of Saccharomyces cerevisiae is important for chromosome transmission and normal cell cycle progression in G2/M. EMBO J. 1990, 9, 4347–4358. [Google Scholar] [CrossRef] [PubMed]
- Parish, J.L.; Rosa, J.; Wang, X.; Lahti, J.M.; Doxsey, S.J.; Androphy, E.J. The DNA helicase ChlR1 is required for sister chromatid cohesion in mammalian cells. J. Cell Sci. 2006, 119, 4857–4865. [Google Scholar] [CrossRef] [PubMed]
- Calì, F.; Bharti, S.K.; Di Perna, R.; Brosh, R.M.; Pisani, F.M. Tim/Timeless, a member of the replication fork protection complex, operates with the Warsaw breakage syndrome DNA helicase DDX11 in the same fork recovery pathway. Nucleic Acids Res. 2016, 44, 705–717. [Google Scholar] [CrossRef] [PubMed]
- Cortone, G.; Zheng, G.; Pensieri, P.; Chiappetta, V.; Tatè, R.; Malacaria, E.; Pichierri, P.; Yu, H.; Pisani, F.M. Interaction of the Warsaw breakage syndrome DNA helicase DDX11 with the replication fork-protection factor Timeless promotes sister chromatid cohesion. PLoS Genet. 2018, 14, e1007622. [Google Scholar] [CrossRef] [PubMed]
- De Lange, T. Shelterin-mediated telomere protection. Annu Rev Genet. 2018, 52, 223–247. [Google Scholar] [CrossRef] [PubMed]
- Porreca, R.M.; Glousker, G.; Awad, A.; Matilla Fernandez, M.I.; Gibaud, A.; Naucke, C.; Cohen, S.B.; Bryan, T.M.; Tzfati, Y.; Draskovic, I.; et al. Human RTEL1 stabilizes long G-overhangs allowing telomerase-dependent over-extension. Nucleic Acids Res. 2018, 46, 4533–4545. [Google Scholar] [CrossRef] [PubMed]
- Uringa, E.J.; Lisaingo, K.; Pickett, H.A.; Brind’Amour, J.; Rohde, J.H.; Zelensky, A.; Essers, J.; Lansdorp, P.M. RTEL1 contributes to DNA replication and repair and telomere maintenance. Mol. Biol. Cell 2012, 23, 2782–2792. [Google Scholar] [CrossRef]
- Vannier, J.B.; Sandhu, S.; Petalcorin, M.I.; Wu, X.; Nabi, Z.; Ding, H.; Boulton, S.J. RTEL1 is a replisome-associated helicase that promotes telomere and genome-wide replication. Science 2013, 342, 239–242. [Google Scholar] [CrossRef]
- Cleaver, J.E.; Lam, E.T.; Revet, I. Disorders of nucleotide excision repair: The genetic and molecular basis of heterogeneity. Nat. Rev. Genet. 2009, 10, 756–768. [Google Scholar] [CrossRef]
- Gray, L.T.; Vallur, A.C.; Eddy, J.; Maizels, N. G quadruplexes are genomewide targets of transcriptional helicases XPB and XPD. Nat. Chem. Biol. 2014, 10, 313–318. [Google Scholar] [CrossRef]
- Tran, H.; Schilling, M.; Wirbelauer, C.; Hess, D.; Nagamine, Y. Facilitation of mRNA deadenylation and decay by the exosome-bound, DExH protein RHAU. Mol. Cell 2004, 13, 101–111. [Google Scholar] [CrossRef]
- Vaughn, J.P.; Creacy, S.D.; Routh, E.D.; Joyner-Butt, C.; Jenkins, G.S.; Pauli, S.; Nagamine, Y.; Akman, S.A. The DEXH protein product of the DHX36 gene is the major source of tetramolecular quadruplex G4-DNA resolving activity in HeLa cell lysates. J. Biol. Chem. 2005, 280, 38117–38120. [Google Scholar] [CrossRef]
- Harrington, C.; Lan, Y.; Akman, S.A. The identification and characterization of a G4-DNA resolvase activity. J. Biol. Chem. 1997, 272, 24631–24636. [Google Scholar] [CrossRef]
- Lai, J.C.; Ponti, S.; Pan, D.; Kohler, H.; Skoda, R.C.; Matthias, P.; Nagamine, Y. The DEAH-box helicase RHAU is an essential gene and critical for mouse hematopoiesis. Blood 2012, 119, 4291–4300. [Google Scholar] [CrossRef] [PubMed]
- Nie, J.; Jiang, M.; Zhang, X.; Tang, H.; Jin, H.; Huang, X.; Yuan, B.; Zhang, C.; Lai, J.C.; Nagamine, Y.; et al. Post-transcriptional Regulation of Nkx2-5 by RHAU in heart development. Cell Rep. 2015, 13, 723–732. [Google Scholar] [CrossRef] [PubMed]
- Creacy, S.D.; Routh, E.D.; Iwamoto, F.; Nagamine, Y.; Akman, S.A.; Vaughn, J.P. G4 resolvase 1 binds both DNA and RNA tetramolecular quadruplex with high affinity and is the major source of tetramolecular quadruplex G4-DNA and G4-RNA resolving activity in HeLa cell lysates. J. Biol. Chem. 2008, 283, 34626–34634. [Google Scholar] [CrossRef] [PubMed]
- Giri, B.; Smaldino, P.J.; Thys, R.G.; Creacy, S.D.; Routh, E.D.; Hantgan, R.R.; Lattmann, S.; Nagamine, Y.; Akman, S.A.; Vaughn, J.P. G4 resolvase 1 tightly binds and unwinds unimolecular G4-DNA. Nucleic Acids Res. 2011, 39, 7161–7178. [Google Scholar] [CrossRef] [PubMed]
- Lattmann, S.; Giri, B.; Vaughn, J.P.; Akman, S.A.; Nagamine, Y. Role of the amino terminal RHAU-specific motif in the recognition and resolution of guanine quadruplex-RNA by the DEAH-box RNA helicase RHAU. Nucleic Acids Res. 2010, 38, 6219–6233. [Google Scholar] [CrossRef]
- Chen, M.C.; Murat, P.; Abecassis, K.; Ferré-D’Amaré, A.R.; Balasubramanian, S. Insights into the mechanism of a G-quadruplex-unwinding DEAH-box helicase. Nucleic Acids Res. 2015, 43, 2223–2231. [Google Scholar] [CrossRef] [PubMed]
- Gueddouda, N.M.; Mendoza, O.; Gomez, D.; Bourdoncle, A.; Mergny, J.L. G-quadruplexes unfolding by RHAU helicase. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 1382–1388. [Google Scholar] [CrossRef]
- Heddi, B.; Cheong, V.V.; Martadinata, H.; Phan, A.T. Insights into G-quadruplex specific recognition by the DEAH-box helicase RHAU: Solution structure of a peptide-quadruplex complex. Proc. Natl. Acad. Sci. USA 2015, 112, 9608–9613. [Google Scholar] [CrossRef] [PubMed]
- Tippana, R.; Xiao, W.; Myong, S. G-quadruplex conformation and dynamics are determined by loop length and sequence. Nucleic Acids Res. 2014, 42, 8106–8114. [Google Scholar] [CrossRef] [PubMed]
- Tippana, R.; Hwang, H.; Opresko, P.L.; Bohr, V.A.; Myong, S. Single-molecule imaging reveals a common mechanism shared by G-quadruplex-resolving helicases. Proc. Natl. Acad. Sci. USA 2016, 113, 8448–8453. [Google Scholar] [CrossRef] [PubMed]
- You, H.; Lattmann, S.; Rhodes, D.; Yan, J. RHAU helicase stabilizes G4 in its nucleotide-free state and destabilizes G4 upon ATP hydrolysis. Nucleic Acids Res. 2017, 45, 206–214. [Google Scholar] [CrossRef] [PubMed]
- Chakraborty, P.; Grosse, F. Human DHX9 helicase preferentially unwinds RNA-containing displacement loops (R-loops) and G-quadruplexes. DNA Repair (Amst.) 2011, 10, 654–665. [Google Scholar] [CrossRef] [PubMed]
- Ohta, S.; Shiomi, Y.; Sugimoto, K.; Obuse, C.; Tsurimoto, T. A proteomics approach to identify proliferating cell nuclear antigen (PCNA)-binding proteins in human cell lysates. Identification of the human CHL12/RFCs2-5 complex as a novel PCNA-binding protein. J. Biol. Chem. 2002, 277, 40362–40367. [Google Scholar] [CrossRef] [PubMed]
- Zhou, K. RNA helicase A interacts with dsDNA and topoisomerase IIα. Nucleic Acids Res. 2003, 31, 2253–2260. [Google Scholar] [CrossRef] [PubMed]
- Friedemann, J.; Grosse, F.; Zhang, S. Nuclear DNA helicase II (RNA helicase A) interacts with Werner syndrome helicase and stimulates its exonuclease activity. J. Biol. Chem. 2005, 280, 31303–31313. [Google Scholar] [CrossRef] [PubMed]
- Lee, T.; Di Paola, D.; Malina, A.; Mills, J.R.; Kreps, A.; Grosse, F.; Tang, H.; Zannis-Hadjopoulos, M.; Larsson, O.; Pelletier, J. Suppression of the DHX9 helicase induces premature senescence in human diploid fibroblasts in a p53-dependent manner. J. Biol. Chem. 2014, 289, 22798–22814. [Google Scholar] [CrossRef] [PubMed]
- Sale, J.E. Competition, collaboration and coordination—Determining how cells bypass DNA damage. J. Cell Sci. 2012, 125, 1633–1643. [Google Scholar] [CrossRef]
- Sale, J.E.; Lehmann, A.R.; Woodgate, R. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nat. Rev. Mol. Cell. Biol. 2012, 13, 141–152. [Google Scholar] [CrossRef] [PubMed]
- Jarmoskaite, I.; Russell, R. RNA helicase proteins as chaperones and remodelers. Annu. Rev. Biochem. 2014, 83, 697–725. [Google Scholar] [CrossRef] [PubMed]
Superfamily | Subfamily | Name | Substrate * | Polarity | Other Functions ¶ | Interactions # | Human Syndrome |
---|---|---|---|---|---|---|---|
SF1 | PIF1 | DNA (parallel and | 5′–3′ | Maintenance of | PCNA | L319P mutation linked to | |
antiparallel, intramolecular ¶ and tetramolecular G4) | mitochondrial genome, DSB ¶ repair | familial breast cancer | |||||
DNA2 | DNA (telomeric G4) | 5′–3′ | Okazaki fragment | RPA | |||
maturation, ICL repair | |||||||
SF2 | Fe-S | FANCJ | DNA (parallel, intramolecular and tetramolecular G4, D- | 5′–3′ | ICL repair, checkpoint control, replication stress | RPA, WRN, BLM, REV1 | Fanconi Anaemia (FA), breast cancer |
loops) | response | ||||||
DDX11 | DNA (antiparallel, | 5′–3′ | Sister chromatid cohesion, | PCNA, RPA, FEN1, | Warsaw Breakage Syndrome (WABS) | ||
intramolecular G4, triplex, 5′ | post-replicative repair | Ctf18-RFC, | |||||
flap, D-loop) | Timeless-Tipin | ||||||
RTEL1 | DNA (telomeric G4) | 5′–3′ | ICL repair, replication stress | POT1, PCNA | Dyskeratosis congenita | ||
response | |||||||
XPD | DNA (parallel tetramolecular | 5′–3′ | NER, basal transcription | Xeroderma Pigmentosum, | |||
G4) | Cockayne syndrome, Trichothiodystrophy (TTD), ¶ Cerebro-oculo-facio-skeletal syndrome (COFS) | ||||||
RecQ | WRN | DNA (intermolecular parallel | 3′–5′ | Fork protection and | RPA, pol δ, FANCJ, | Werner syndrome | |
G4, triplex, Holliday Junction) | remodelling, replication | POT1, BLM | |||||
stress response, checkpoint | |||||||
control, DSB and ICL repair | |||||||
BLM | DNA (intermolecular and | 3′–5′ | Fork protection and | RPA, pol δ, FANCJ, | Bloom syndrome | ||
tetramolecular parallel G4, | remodelling, DSB and ICL | POT1, WRN | |||||
triplex, Holliday Junction) | repair | ||||||
DEAH | DHX9 | RNA and DNA (intramolecular and tetramolecular parallel | 3′–5′ | Replication stress response, transcription regulation | PCNA, TOP2α, WRN | ||
G4, R-loop) | |||||||
RHAU/DHX36 | RNA and DNA (intramolecular | 3′–5′ | Translation regulation, RNA | ||||
and tetramolecular parallel ζ G4, R-loop) | decay, telomere length ζ regulation |
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).
Share and Cite
Lerner, L.K.; Sale, J.E. Replication of G Quadruplex DNA. Genes 2019, 10, 95. https://doi.org/10.3390/genes10020095
Lerner LK, Sale JE. Replication of G Quadruplex DNA. Genes. 2019; 10(2):95. https://doi.org/10.3390/genes10020095
Chicago/Turabian StyleLerner, Leticia Koch, and Julian E. Sale. 2019. "Replication of G Quadruplex DNA" Genes 10, no. 2: 95. https://doi.org/10.3390/genes10020095
APA StyleLerner, L. K., & Sale, J. E. (2019). Replication of G Quadruplex DNA. Genes, 10(2), 95. https://doi.org/10.3390/genes10020095