Common Chemical Inductors of Replication Stress: Focus on Cell-Based Studies
Language English Country Switzerland Media electronic
Document type Journal Article, Review, Research Support, Non-U.S. Gov't
PubMed
28230817
PubMed Central
PMC5372731
DOI
10.3390/biom7010019
PII: biom7010019
Knihovny.cz E-resources
- Keywords
- replication stress, aphidicolin, camptothecin, cancer, cisplatin, etoposide, hydroxyurea,
- MeSH
- Cells drug effects metabolism MeSH
- Stress, Physiological * drug effects MeSH
- Humans MeSH
- Antineoplastic Agents chemistry pharmacology MeSH
- DNA Replication * drug effects MeSH
- Animals MeSH
- Check Tag
- Humans MeSH
- Animals MeSH
- Publication type
- Journal Article MeSH
- Research Support, Non-U.S. Gov't MeSH
- Review MeSH
- Names of Substances
- Antineoplastic Agents MeSH
DNA replication is a highly demanding process regarding the energy and material supply and must be precisely regulated, involving multiple cellular feedbacks. The slowing down or stalling of DNA synthesis and/or replication forks is referred to as replication stress (RS). Owing to the complexity and requirements of replication, a plethora of factors may interfere and challenge the genome stability, cell survival or affect the whole organism. This review outlines chemical compounds that are known inducers of RS and commonly used in laboratory research. These compounds act on replication by direct interaction with DNA causing DNA crosslinks and bulky lesions (cisplatin), chemical interference with the metabolism of deoxyribonucleotide triphosphates (hydroxyurea), direct inhibition of the activity of replicative DNA polymerases (aphidicolin) and interference with enzymes dealing with topological DNA stress (camptothecin, etoposide). As a variety of mechanisms can induce RS, the responses of mammalian cells also vary. Here, we review the activity and mechanism of action of these compounds based on recent knowledge, accompanied by examples of induced phenotypes, cellular readouts and commonly used doses.
See more in PubMed
Zeman M.K., Cimprich K.A. Causes and consequences of replication stress. Nat. Cell Biol. 2014;16:2–9. doi: 10.1038/ncb2897. PubMed DOI PMC
Burhans W.C., Weinberger M. DNA replication stress, genome instability and aging. Nucleic Acids Res. 2007;35:7545–7556. doi: 10.1093/nar/gkm1059. PubMed DOI PMC
Huh M.S., Ivanochko D., Hashem L.E., Curtin M., Delorme M., Goodall E., Yan K., Picketts D.J. Stalled replication forks within heterochromatin require ATRX for protection. Cell Death Dis. 2016;7:e2220. doi: 10.1038/cddis.2016.121. PubMed DOI PMC
Gelot C., Magdalou I., Lopez B.S. Replication stress in Mammalian cells and its consequences for mitosis. Genes. 2015;6:267–298. doi: 10.3390/genes6020267. PubMed DOI PMC
Krasilnikova M.M., Mirkin S.M. Replication stalling at Friedreich’s ataxia (GAA)n repeats in vivo. Mol. Cell. Biol. 2004;24:2286–2295. doi: 10.1128/MCB.24.6.2286-2295.2004. PubMed DOI PMC
Neelsen K.J., Zanini I.M.Y., Mijic S., Herrador R., Zellweger R., Ray Chaudhuri A., Creavin K.D., Blow J.J., Lopes M. Deregulated origin licensing leads to chromosomal breaks by rereplication of a gapped DNA template. Genes Dev. 2013;27:2537–2542. doi: 10.1101/gad.226373.113. PubMed DOI PMC
Porter A.C. Preventing DNA over-replication: A Cdk perspective. Cell Div. 2008;3:3. doi: 10.1186/1747-1028-3-3. PubMed DOI PMC
Burrell R.A., McClelland S.E., Endesfelder D., Groth P., Weller M.-C., Shaikh N., Domingo E., Kanu N., Dewhurst S.M., Gronroos E., et al. Replication stress links structural and numerical cancer chromosomal instability. Nature. 2013;494:492–496. doi: 10.1038/nature11935. PubMed DOI PMC
Liu B., Alberts B.M. Head-on collision between a DNA replication apparatus and RNA polymerase transcription complex. Science. 1995;267:1131–1137. doi: 10.1126/science.7855590. PubMed DOI
Bartkova J., Rezaei N., Liontos M., Karakaidos P., Kletsas D., Issaeva N., Vassiliou L.-V.F., Kolettas E., Niforou K., Zoumpourlis V.C., et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature. 2006;444:633–637. doi: 10.1038/nature05268. PubMed DOI
Vallerga M.B., Mansilla S.F., Federico M.B., Bertolin A.P., Gottifredi V. Rad51 recombinase prevents Mre11 nuclease-dependent degradation and excessive PrimPol-mediated elongation of nascent DNA after UV irradiation. Proc. Natl. Acad. Sci. USA. 2015;112:E6624–E6633. doi: 10.1073/pnas.1508543112. PubMed DOI PMC
Mazouzi A., Velimezi G., Loizou J.I. DNA replication stress: Causes, resolution and disease. Exp. Cell Res. 2014;329:85–93. doi: 10.1016/j.yexcr.2014.09.030. PubMed DOI
Jekimovs C., Bolderson E., Suraweera A., Adams M., O’Byrne K.J., Richard D.J. Chemotherapeutic compounds targeting the DNA double-strand break repair pathways: The good, the bad, and the promising. Front. Oncol. 2014;4:86. doi: 10.3389/fonc.2014.00086. PubMed DOI PMC
Beranek D.T. Distribution of methyl and ethyl adducts following alkylation with monofunctional alkylating agents. Mutat. Res. 1990;231:11–30. doi: 10.1016/0027-5107(90)90173-2. PubMed DOI
Kondo N., Takahashi A., Mori E., Noda T., Su X., Ohnishi K., McKinnon P.J., Sakaki T., Nakase H., Ono K., et al. DNA ligase IV is a potential molecular target in ACNU sensitivity. Cancer Sci. 2010;101:1881–1885. doi: 10.1111/j.1349-7006.2010.01591.x. PubMed DOI PMC
Brookes P., Lawley P.D. The reaction of mono- and di-functional alkylating agents with nucleic acids. Biochem. J. 1961;80:496–503. doi: 10.1042/bj0800496. PubMed DOI PMC
Lawley P.D., Brookes P. The action of alkylating agents on deoxyribonucleic acid in relation to biological effects of the alkylating agents. Exp. Cell Res. 1963;24(Suppl. S9):512–520. doi: 10.1016/0014-4827(63)90291-X. PubMed DOI
Noll D.M., Mason T.M., Miller P.S. Formation and repair of interstrand cross-links in DNA. Chem. Rev. 2006;106:277–301. doi: 10.1021/cr040478b. PubMed DOI PMC
Schärer O.D. DNA Interstrand Crosslinks: Natural and Drug-Induced DNA Adducts that Induce Unique Cellular Responses. ChemBioChem. 2005;6:27–32. doi: 10.1002/cbic.200400287. PubMed DOI
Lawley P.D., Phillips D.H. DNA adducts from chemotherapeutic agents. Mutat. Res. 1996;355:13–40. doi: 10.1016/0027-5107(96)00020-6. PubMed DOI
Bhuyan B.K., Scheidt L.G., Fraser T.J. Cell cycle phase specificity of antitumor agents. Cancer Res. 1972;32:398–407. PubMed
Glover T.W., Arlt M.F., Casper A.M., Durkin S.G. Mechanisms of common fragile site instability. Hum. Mol. Genet. 2005;14:R197–R205. doi: 10.1093/hmg/ddi265. PubMed DOI
Koç A., Wheeler L.J., Mathews C.K., Merrill G.F. Hydroxyurea arrests DNA replication by a mechanism that preserves basal dNTP pools. J. Biol. Chem. 2004;279:223–230. doi: 10.1074/jbc.M303952200. PubMed DOI
Hsiang Y.H., Lihou M.G., Liu L.F. Arrest of replication forks by drug-stabilized topoisomerase I-DNA cleavable complexes as a mechanism of cell killing by camptothecin. Cancer Res. 1989;49:5077–5082. PubMed
Deweese J.E., Osheroff N. The DNA cleavage reaction of topoisomerase II: Wolf in sheep’s clothing. Nucleic Acids Res. 2009;37:738–748. doi: 10.1093/nar/gkn937. PubMed DOI PMC
Helleday T., Petermann E., Lundin C., Hodgson B., Sharma R.A. DNA repair pathways as targets for cancer therapy. Nat. Rev. Cancer. 2008;8:193–204. doi: 10.1038/nrc2342. PubMed DOI
Krokan H.E., Bjørås M. Base Excision Repair. Cold Spring Harb. Perspect. Biol. 2013;5:a012583. doi: 10.1101/cshperspect.a012583. PubMed DOI PMC
Gillet L.C.J., Schärer O.D. Molecular mechanisms of mammalian global genome nucleotide excision repair. Chem. Rev. 2006;106:253–276. doi: 10.1021/cr040483f. PubMed DOI
Caldecott K.W. Protein ADP-ribosylation and the cellular response to DNA strand breaks. DNA Repair. 2014;19:108–113. doi: 10.1016/j.dnarep.2014.03.021. PubMed DOI
Heyer W.-D., Ehmsen K.T., Liu J. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 2010;44:113–139. doi: 10.1146/annurev-genet-051710-150955. PubMed DOI PMC
Davis A.J., Chen D.J. DNA double strand break repair via non-homologous end-joining. Transl. Cancer Res. 2013;2:130–143. PubMed PMC
Bi X. Mechanism of DNA damage tolerance. World J. Biol. Chem. 2015;6:48–56. doi: 10.4331/wjbc.v6.i3.48. PubMed DOI PMC
Aguilera A., Gómez-González B. Genome instability: A mechanistic view of its causes and consequences. Nat. Rev. Genet. 2008;9:204–217. doi: 10.1038/nrg2268. PubMed DOI
Chang D.J., Cimprich K.A. DNA damage tolerance: When it’s OK to make mistakes. Nat. Chem. Biol. 2009;5:82–90. doi: 10.1038/nchembio.139. PubMed DOI PMC
Ghosal G., Chen J. DNA damage tolerance: A double-edged sword guarding the genome. Transl. Cancer Res. 2013;2:107–129. PubMed PMC
Saugar I., Ortiz-Bazán M.Á., Tercero J.A. Tolerating DNA damage during eukaryotic chromosome replication. Exp. Cell Res. 2014;329:170–177. doi: 10.1016/j.yexcr.2014.07.009. PubMed DOI
Deans A.J., West S.C. DNA interstrand crosslink repair and cancer. Nat. Rev. Cancer. 2011;11:467–480. doi: 10.1038/nrc3088. PubMed DOI PMC
Longerich S., Li J., Xiong Y., Sung P., Kupfer G.M. Stress and DNA repair biology of the Fanconi anemia pathway. Blood. 2014;124:2812–2819. doi: 10.1182/blood-2014-04-526293. PubMed DOI PMC
Gaillard H., García-Muse T., Aguilera A. Replication stress and cancer. Nat. Rev. Cancer. 2015;15:276–289. doi: 10.1038/nrc3916. PubMed DOI
Mamrak N.E., Shimamura A., Howlett N.G. Recent discoveries in the molecular pathogenesis of the inherited bone marrow failure syndrome Fanconi anemia. Blood Rev. 2016 doi: 10.1016/j.blre.2016.10.002. PubMed DOI PMC
Kennedy R.D., D’Andrea A.D. The Fanconi Anemia/BRCA pathway: New faces in the crowd. Genes Dev. 2005;19:2925–2940. doi: 10.1101/gad.1370505. PubMed DOI
Thompson L.H., Hinz J.M. Cellular and molecular consequences of defective Fanconi anemia proteins in replication-coupled DNA repair: Mechanistic insights. Mutat. Res. 2009;668:54–72. doi: 10.1016/j.mrfmmm.2009.02.003. PubMed DOI PMC
Branzei D. Ubiquitin family modifications and template switching. FEBS Lett. 2011;585:2810–2817. doi: 10.1016/j.febslet.2011.04.053. PubMed DOI
Xu X., Blackwell S., Lin A., Li F., Qin Z., Xiao W. Error-free DNA-damage tolerance in Saccharomyces cerevisiae. Mutat. Res. Rev. Mutat. Res. 2015;764:43–50. doi: 10.1016/j.mrrev.2015.02.001. PubMed DOI
Ge X.Q., Jackson D.A., Blow J.J. Dormant origins licensed by excess Mcm2–7 are required for human cells to survive replicative stress. Genes Dev. 2007;21:3331–3341. doi: 10.1101/gad.457807. PubMed DOI PMC
Lopes M., Foiani M., Sogo J.M. Multiple mechanisms control chromosome integrity after replication fork uncoupling and restart at irreparable UV lesions. Mol. Cell. 2006;21:15–27. doi: 10.1016/j.molcel.2005.11.015. PubMed DOI
Woodward A.M., Göhler T., Luciani M.G., Oehlmann M., Ge X., Gartner A., Jackson D.A., Blow J.J. Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress. J. Cell Biol. 2006;173:673–683. doi: 10.1083/jcb.200602108. PubMed DOI PMC
Elvers I., Johansson F., Groth P., Erixon K., Helleday T. UV stalled replication forks restart by re-priming in human fibroblasts. Nucleic Acids Res. 2011;39:7049–7057. doi: 10.1093/nar/gkr420. PubMed DOI PMC
McIntosh D., Blow J.J. Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb. Perspect. Biol. 2012;4:a012955. doi: 10.1101/cshperspect.a012955. PubMed DOI PMC
De Piccoli G., Katou Y., Itoh T., Nakato R., Shirahige K., Labib K. Replisome stability at defective DNA replication forks is independent of S phase checkpoint kinases. Mol. Cell. 2012;45:696–704. doi: 10.1016/j.molcel.2012.01.007. PubMed DOI
Tercero J.A., Diffley J.F.X. Regulation of DNA replication fork progression through damaged DNA by the Mec1/Rad53 checkpoint. Nature. 2001;412:553–557. doi: 10.1038/35087607. PubMed DOI
Lopes M., Cotta-Ramusino C., Pellicioli A., Liberi G., Plevani P., Muzi-Falconi M., Newlon C.S., Foiani M. The DNA replication checkpoint response stabilizes stalled replication forks. Nature. 2001;412:557–561. doi: 10.1038/35087613. PubMed DOI
Cobb J.A., Bjergbaek L., Shimada K., Frei C., Gasser S.M. DNA polymerase stabilization at stalled replication forks requires Mec1 and the RecQ helicase Sgs1. EMBO J. 2003;22:4325–4336. doi: 10.1093/emboj/cdg391. PubMed DOI PMC
Ragland R.L., Patel S., Rivard R.S., Smith K., Peters A.A., Bielinsky A.-K., Brown E.J. RNF4 and PLK1 are required for replication fork collapse in ATR-deficient cells. Genes Dev. 2013;27:2259–2273. doi: 10.1101/gad.223180.113. PubMed DOI PMC
Hanada K., Budzowska M., Davies S.L., van Drunen E., Onizawa H., Beverloo H.B., Maas A., Essers J., Hickson I.D., Kanaar R. The structure-specific endonuclease Mus81 contributes to replication restart by generating double-strand DNA breaks. Nat. Struct. Mol. Biol. 2007;14:1096–1104. doi: 10.1038/nsmb1313. PubMed DOI
Forment J.V., Blasius M., Guerini I., Jackson S.P. Structure-specific DNA endonuclease Mus81/Eme1 generates DNA damage caused by Chk1 inactivation. PLoS ONE. 2011;6:e23517. doi: 10.1371/journal.pone.0023517. PubMed DOI PMC
Zellweger R., Dalcher D., Mutreja K., Berti M., Schmid J.A., Herrador R., Vindigni A., Lopes M. Rad51-mediated replication fork reversal is a global response to genotoxic treatments in human cells. J. Cell Biol. 2015;208:563–579. doi: 10.1083/jcb.201406099. PubMed DOI PMC
Pacek M., Walter J.C. A requirement for MCM7 and Cdc45 in chromosome unwinding during eukaryotic DNA replication. EMBO J. 2004;23:3667–3676. doi: 10.1038/sj.emboj.7600369. PubMed DOI PMC
Byun T.S., Pacek M., Yee M., 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. doi: 10.1101/gad.1301205. PubMed DOI PMC
Zou L., Elledge S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science. 2003;300:1542–1548. doi: 10.1126/science.1083430. PubMed DOI
MacDougall C.A., Byun T.S., Van C., Yee M., Cimprich K.A. The structural determinants of checkpoint activation. Genes Dev. 2007;21:898–903. doi: 10.1101/gad.1522607. PubMed DOI PMC
Nam E.A., Cortez D. ATR signalling: More than meeting at the fork. Biochem. J. 2011;436:527–536. doi: 10.1042/BJ20102162. PubMed DOI PMC
Maréchal A., Zou L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 2013;5:a012716. doi: 10.1101/cshperspect.a012716. PubMed DOI PMC
Lucca C., Vanoli F., Cotta-Ramusino C., Pellicioli A., Liberi G., Haber J., Foiani M. Checkpoint-mediated control of replisome-fork association and signalling in response to replication pausing. Oncogene. 2004;23:1206–1213. doi: 10.1038/sj.onc.1207199. PubMed DOI
Petermann E., Orta M.L., Issaeva N., Schultz N., Helleday T. Hydroxyurea-stalled replication forks become progressively inactivated and require two different RAD51-mediated pathways for restart and repair. Mol. Cell. 2010;37:492–502. doi: 10.1016/j.molcel.2010.01.021. PubMed DOI PMC
Labib K., De Piccoli G. Surviving chromosome replication: The many roles of the S-phase checkpoint pathway. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2011;366:3554–3561. doi: 10.1098/rstb.2011.0071. PubMed DOI PMC
Ozeri-Galai E., Schwartz M., Rahat A., Kerem B. Interplay between ATM and ATR in the regulation of common fragile site stability. Oncogene. 2008;27:2109–2117. doi: 10.1038/sj.onc.1210849. PubMed DOI
Ciccia A., Elledge S.J. The DNA damage response: Making it safe to play with knives. Mol. Cell. 2010;40:179–204. doi: 10.1016/j.molcel.2010.09.019. PubMed DOI PMC
Ammazzalorso F., Pirzio L.M., Bignami M., Franchitto A., Pichierri P. ATR and ATM differently regulate WRN to prevent DSBs at stalled replication forks and promote replication fork recovery. EMBO J. 2010;29:3156–3169. doi: 10.1038/emboj.2010.205. PubMed DOI PMC
Bachrati C.Z., Hickson I.D. RecQ helicases: Suppressors of tumorigenesis and premature aging. Biochem. J. 2003;374:577–606. doi: 10.1042/bj20030491. PubMed DOI PMC
Hills S.A., Diffley J.F.X. DNA replication and oncogene-induced replicative stress. Curr. Biol. 2014;24:R435–R444. doi: 10.1016/j.cub.2014.04.012. PubMed DOI
Macheret M., Halazonetis T.D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 2015;10:425–448. doi: 10.1146/annurev-pathol-012414-040424. PubMed DOI
Murga M., Campaner S., Lopez-Contreras A.J., Toledo L.I., Soria R., Montaña M.F., D’Artista L., Schleker T., Guerra C., Garcia E., et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat. Struct. Mol. Biol. 2011;18:1331–1335. doi: 10.1038/nsmb.2189. PubMed DOI PMC
Marusyk A., Wheeler L.J., Mathews C.K., DeGregori J. p53 mediates senescence-like arrest induced by chronic replicational stress. Mol. Cell. Biol. 2007;27:5336–5351. doi: 10.1128/MCB.01316-06. PubMed DOI PMC
Bai G., Smolka M.B., Schimenti J.C. Chronic DNA Replication Stress Reduces Replicative Lifespan of Cells by TRP53-Dependent, microRNA-Assisted MCM2–7 Downregulation. PLoS Genet. 2016;12:e1005787. doi: 10.1371/journal.pgen.1005787. PubMed DOI PMC
Bartkova J., Horejsí Z., Koed K., Krämer A., Tort F., Zieger K., Guldberg P., Sehested M., Nesland J.M., Lukas C., et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature. 2005;434:864–870. doi: 10.1038/nature03482. PubMed DOI
O’Driscoll M., Ruiz-Perez V.L., Woods C.G., Jeggo P.A., Goodship J.A. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nat. Genet. 2003;33:497–501. doi: 10.1038/ng1129. PubMed DOI
McKinnon P.J. ATM and ataxia telangiectasia. EMBO Rep. 2004;5:772–776. doi: 10.1038/sj.embor.7400210. PubMed DOI PMC
DiGiovanna J.J., Kraemer K.H. Shining a light on xeroderma pigmentosum. J. Investig. Dermatol. 2012;132:785–796. doi: 10.1038/jid.2011.426. PubMed DOI PMC
Callén E., Surrallés J. Telomere dysfunction in genome instability syndromes. Mutat. Res. 2004;567:85–104. doi: 10.1016/j.mrrev.2004.06.003. PubMed DOI
Lauper J.M., Krause A., Vaughan T.L., Monnat R.J. Spectrum and risk of neoplasia in Werner syndrome: A systematic review. PLoS ONE. 2013;8:e59709. doi: 10.1371/journal.pone.0059709. PubMed DOI PMC
Bernstein K.A., Gangloff S., Rothstein R. The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 2010;44:393–417. doi: 10.1146/annurev-genet-102209-163602. PubMed DOI PMC
Kim H., D’Andrea A.D. Regulation of DNA cross-link repair by the Fanconi anemia/BRCA pathway. Genes Dev. 2012;26:1393–1408. doi: 10.1101/gad.195248.112. PubMed DOI PMC
Joenje H., Patel K.J. The emerging genetic and molecular basis of Fanconi anaemia. Nat. Rev. Genet. 2001;2:446–457. doi: 10.1038/35076590. PubMed DOI
Larizza L., Roversi G., Volpi L. Rothmund-Thomson syndrome. Orphanet J. Rare Dis. 2010;5:2. doi: 10.1186/1750-1172-5-2. PubMed DOI PMC
Lu H., Shamanna R.A., Keijzers G., Anand R., Rasmussen L.J., Cejka P., Croteau D.L., Bohr V.A. RECQL4 Promotes DNA End Resection in Repair of DNA Double-Strand Breaks. Cell Rep. 2016;16:161–173. doi: 10.1016/j.celrep.2016.05.079. PubMed DOI PMC
Rosenberg B., Vancamp L., Krigas T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature. 1965;205:698–699. doi: 10.1038/205698a0. PubMed DOI
Todd R.C., Lippard S.J. Inhibition of transcription by platinum antitumor compounds. Metallomics. 2009;1:280–291. doi: 10.1039/b907567d. PubMed DOI PMC
Zamble D.B., Lippard S.J. Cisplatin and DNA repair in cancer chemotherapy. Trends Biochem. Sci. 1995;20:435–439. doi: 10.1016/S0968-0004(00)89095-7. PubMed DOI
[(accessed on 23 January 2017)]. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=3CO3.
Fichtinger-Schepman A.M., van der Veer J.L., den Hartog J.H., Lohman P.H., Reedijk J. Adducts of the antitumor drug cis-diamminedichloroplatinum(II) with DNA: Formation, identification, and quantitation. Biochemistry. 1985;24:707–713. doi: 10.1021/bi00324a025. PubMed DOI
Harder H.C., Rosenberg B. Inhibitory effects of anti-tumor platinum compounds on DNA, RNA and protein syntheses in mammalian cells in virtro. Int. J. Cancer. 1970;6:207–216. doi: 10.1002/ijc.2910060207. PubMed DOI
Eastman A. Reevaluation of interaction of cis-dichloro(ethylenediamine)platinum(II) with DNA. Biochemistry. 1986;25:3912–3915. doi: 10.1021/bi00361a026. PubMed DOI
Sherman S.E., Gibson D., Wang A.H., Lippard S.J. X-ray structure of the major adduct of the anticancer drug cisplatin with DNA: cis-[Pt(NH3)2(d(pGpG))] Science. 1985;230:412–417. doi: 10.1126/science.4048939. PubMed DOI
Eastman A. Separation and characterization of products resulting from the reaction of cis-diamminedichloroplatinum (II) with deoxyribonucleosides. Biochemistry. 1982;21:6732–6736. doi: 10.1021/bi00269a018. PubMed DOI
Desoize B. Cancer and metals and metal compounds: Part I—Carcinogenesis. Crit. Rev. Oncol. Hematol. 2002;42:1–3. doi: 10.1016/S1040-8428(02)00017-3. PubMed DOI
Köberle B., Masters J.R., Hartley J.A., Wood R.D. Defective repair of cisplatin-induced DNA damage caused by reduced XPA protein in testicular germ cell tumours. Curr. Biol. 1999;9:273–276. doi: 10.1016/S0960-9822(99)80118-3. PubMed DOI
Borst P., Rottenberg S., Jonkers J. How do real tumors become resistant to cisplatin? Cell Cycle. 2008;7:1353–1359. doi: 10.4161/cc.7.10.5930. PubMed DOI
Sedletska Y., Fourrier L., Malinge J.-M. Modulation of MutS ATP-dependent functional activities by DNA containing a cisplatin compound lesion (base damage and mismatch) J. Mol. Biol. 2007;369:27–40. doi: 10.1016/j.jmb.2007.02.048. PubMed DOI
Brown R., Clugston C., Burns P., Edlin A., Vasey P., Vojtĕsek B., Kaye S.B. Increased accumulation of p53 protein in cisplatin-resistant ovarian cell lines. Int. J. Cancer. 1993;55:678–684. doi: 10.1002/ijc.2910550428. PubMed DOI
Damsma G.E., Alt A., Brueckner F., Carell T., Cramer P. Mechanism of transcriptional stalling at cisplatin-damaged DNA. Nat. Struct. Mol. Biol. 2007;14:1127–1133. doi: 10.1038/nsmb1314. PubMed DOI
Shimodaira H., Yoshioka-Yamashita A., Kolodner R.D., Wang J.Y.J. Interaction of mismatch repair protein PMS2 and the p53-related transcription factor p73 in apoptosis response to cisplatin. Proc. Natl. Acad. Sci. USA. 2003;100:2420–2425. doi: 10.1073/pnas.0438031100. PubMed DOI PMC
Aebi S., Kurdi-Haidar B., Gordon R., Cenni B., Zheng H., Fink D., Christen R.D., Boland C.R., Koi M., Fishel R., et al. Loss of DNA mismatch repair in acquired resistance to cisplatin. Cancer Res. 1996;56:3087–3090. PubMed
Alt A., Lammens K., Chiocchini C., Lammens A., Pieck J.C., Kuch D., Hopfner K.-P., Carell T. Bypass of DNA lesions generated during anticancer treatment with cisplatin by DNA polymerase eta. Science. 2007;318:967–970. doi: 10.1126/science.1148242. PubMed DOI
Brozovic A., Ambriović-Ristov A., Osmak M. The relationship between cisplatin-induced reactive oxygen species, glutathione, and BCL-2 and resistance to cisplatin. Crit. Rev. Toxicol. 2010;40:347–359. doi: 10.3109/10408441003601836. PubMed DOI
Splettstoesser F., Florea A.-M., Büsselberg D. IP3 receptor antagonist, 2-APB, attenuates cisplatin induced Ca2+-influx in HeLa-S3 cells and prevents activation of calpain and induction of apoptosis. Br. J. Pharmacol. 2007;151:1176–1186. doi: 10.1038/sj.bjp.0707335. PubMed DOI PMC
Shamimi-Noori S., Yeow W.-S., Ziauddin M.F., Xin H., Tran T.L.N., Xie J., Loehfelm A., Patel P., Yang J., Schrump D.S., et al. Cisplatin enhances the antitumor effect of tumor necrosis factor-related apoptosis-inducing ligand gene therapy via recruitment of the mitochondria-dependent death signaling pathway. Cancer Gene Ther. 2008;15:356–370. doi: 10.1038/sj.cgt.7701120. PubMed DOI
Qian W., Nishikawa M., Haque A.M., Hirose M., Mashimo M., Sato E., Inoue M. Mitochondrial density determines the cellular sensitivity to cisplatin-induced cell death. Am. J. Physiol. Cell Physiol. 2005;289:C1466–C1475. doi: 10.1152/ajpcell.00265.2005. PubMed DOI
Wetzel C.C., Berberich S.J. p53 binds to cisplatin-damaged DNA. Biochim. Biophys. Acta. 2001;1517:392–397. doi: 10.1016/S0167-4781(00)00305-5. PubMed DOI
Kutuk O., Arisan E.D., Tezil T., Shoshan M.C., Basaga H. Cisplatin overcomes Bcl-2-mediated resistance to apoptosis via preferential engagement of Bak: Critical role of Noxa-mediated lipid peroxidation. Carcinogenesis. 2009;30:1517–1527. doi: 10.1093/carcin/bgp165. PubMed DOI
Kim H.-S., Hwang J.-T., Yun H., Chi S.-G., Lee S.-J., Kang I., Yoon K.-S., Choe W.-J., Kim S.-S., Ha J. Inhibition of AMP-activated protein kinase sensitizes cancer cells to cisplatin-induced apoptosis via hyper-induction of p53. J. Biol. Chem. 2008;283:3731–3742. doi: 10.1074/jbc.M704432200. PubMed DOI
Yang C., Kaushal V., Haun R.S., Seth R., Shah S.V., Kaushal G.P. Transcriptional activation of caspase-6 and -7 genes by cisplatin-induced p53 and its functional significance in cisplatin nephrotoxicity. Cell Death Differ. 2008;15:530–544. doi: 10.1038/sj.cdd.4402287. PubMed DOI
Jiang M., Wei Q., Wang J., Du Q., Yu J., Zhang L., Dong Z. Regulation of PUMA-alpha by p53 in cisplatin-induced renal cell apoptosis. Oncogene. 2006;25:4056–4066. doi: 10.1038/sj.onc.1209440. PubMed DOI
Righetti S.C., Della Torre G., Pilotti S., Ménard S., Ottone F., Colnaghi M.I., Pierotti M.A., Lavarino C., Cornarotti M., Oriana S., et al. A comparative study of p53 gene mutations, protein accumulation, and response to cisplatin-based chemotherapy in advanced ovarian carcinoma. Cancer Res. 1996;56:689–693. PubMed
Johnson C.L., Lu D., Huang J., Basu A. Regulation of p53 stabilization by DNA damage and protein kinase C. Mol. Cancer Ther. 2002;1:861–867. PubMed
Gong J.G., Costanzo A., Yang H.Q., Melino G., Kaelin W.G., Levrero M., Wang J.Y. The tyrosine kinase c-Abl regulates p73 in apoptotic response to cisplatin-induced DNA damage. Nature. 1999;399:806–809. PubMed
Preyer M., Shu C.-W., Wang J.Y.J. Delayed activation of Bax by DNA damage in embryonic stem cells with knock-in mutations of the Abl nuclear localization signals. Cell Death Differ. 2007;14:1139–1148. doi: 10.1038/sj.cdd.4402119. PubMed DOI
Tsai K.K.C., Yuan Z.-M. c-Abl stabilizes p73 by a phosphorylation-augmented interaction. Cancer Res. 2003;63:3418–3424. PubMed
Levy D., Adamovich Y., Reuven N., Shaul Y. Yap1 phosphorylation by c-Abl is a critical step in selective activation of proapoptotic genes in response to DNA damage. Mol. Cell. 2008;29:350–361. doi: 10.1016/j.molcel.2007.12.022. PubMed DOI
Jones E.V., Dickman M.J., Whitmarsh A.J. Regulation of p73-mediated apoptosis by c-Jun N-terminal kinase. Biochem. J. 2007;405:617–623. doi: 10.1042/BJ20061778. PubMed DOI PMC
Hayakawa J., Ohmichi M., Kurachi H., Kanda Y., Hisamoto K., Nishio Y., Adachi K., Tasaka K., Kanzaki T., Murata Y. Inhibition of BAD phosphorylation either at serine 112 via extracellular signal-regulated protein kinase cascade or at serine 136 via Akt cascade sensitizes human ovarian cancer cells to cisplatin. Cancer Res. 2000;60:5988–5994. PubMed
Isonishi S., Andrews P.A., Howell S.B. Increased sensitivity to cis-diamminedichloroplatinum(II) in human ovarian carcinoma cells in response to treatment with 12-O-tetradecanoylphorbol 13-acetate. J. Biol. Chem. 1990;265:3623–3627. PubMed
Basu A., Teicher B.A., Lazo J.S. Involvement of protein kinase C in phorbol ester-induced sensitization of HeLa cells to cis-diamminedichloroplatinum(II) J. Biol. Chem. 1990;265:8451–8457. PubMed
Wang X., Dhalla N.S. Modification of beta-adrenoceptor signal transduction pathway by genetic manipulation and heart failure. Mol. Cell. Biochem. 2000;214:131–155. doi: 10.1023/A:1007131925048. PubMed DOI
Basu A., Tu H. Activation of ERK during DNA damage-induced apoptosis involves protein kinase Cdelta. Biochem. Biophys. Res. Commun. 2005;334:1068–1073. doi: 10.1016/j.bbrc.2005.06.199. PubMed DOI
Nowak G. Protein kinase C-alpha and ERK1/2 mediate mitochondrial dysfunction, decreases in active Na+ transport, and cisplatin-induced apoptosis in renal cells. J. Biol. Chem. 2002;277:43377–43388. doi: 10.1074/jbc.M206373200. PubMed DOI PMC
Sánchez-Pérez I., Benitah S.A., Martínez-Gomariz M., Lacal J.C., Perona R. Cell stress and MEKK1-mediated c-Jun activation modulate NFκB activity and cell viability. Mol. Biol. Cell. 2002;13:2933–2945. doi: 10.1091/mbc.E02-01-0022. PubMed DOI PMC
Jones J.A., Stroud R.E., Kaplan B.S., Leone A.M., Bavaria J.E., Gorman J.H., Gorman R.C., Ikonomidis J.S. Differential protein kinase C isoform abundance in ascending aortic aneurysms from patients with bicuspid versus tricuspid aortic valves. Circulation. 2007;116:I144–I149. doi: 10.1161/CIRCULATIONAHA.106.681361. PubMed DOI
Zanke B.W., Boudreau K., Rubie E., Winnett E., Tibbles L.A., Zon L., Kyriakis J., Liu F.F., Woodgett J.R. The stress-activated protein kinase pathway mediates cell death following injury induced by cis-platinum, UV irradiation or heat. Curr. Biol. 1996;6:606–613. doi: 10.1016/S0960-9822(02)00547-X. PubMed DOI
Hernández Losa J., Parada Cobo C., Guinea Viniegra J., Sánchez-Arevalo Lobo V.J., Ramón y Cajal S., Sánchez-Prieto R. Role of the p38 MAPK pathway in cisplatin-based therapy. Oncogene. 2003;22:3998–4006. doi: 10.1038/sj.onc.1206608. PubMed DOI
Wang S.J., Bourguignon L.Y.W. Hyaluronan-CD44 promotes phospholipase C-mediated Ca2+ signaling and cisplatin resistance in head and neck cancer. Arch. Otolaryngol. Head Neck Surg. 2006;132:19–24. doi: 10.1001/archotol.132.1.19. PubMed DOI
Speelmans G., Staffhorst R.W., Versluis K., Reedijk J., de Kruijff B. Cisplatin complexes with phosphatidylserine in membranes. Biochemistry. 1997;36:10545–10550. doi: 10.1021/bi9703047. PubMed DOI
Huihui Z., Baohuai W., Youming Z., Kui W. Calorimetric studies on actin polymerization and a comparison of the effects of cisplatin and transplatin. Thermochim. Acta. 1995;265:31–38. doi: 10.1016/0040-6031(95)02425-2. DOI
Chen X., Jiang Y., Huang Z., Li D., Chen X., Cao M., Meng Q., Pang H., Sun L., Zhao Y., et al. miRNA-378 reverses chemoresistance to cisplatin in lung adenocarcinoma cells by targeting secreted clusterin. Sci. Rep. 2016;6:19455. doi: 10.1038/srep19455. PubMed DOI PMC
Zhu H., Wu H., Liu X., Evans B.R., Medina D.J., Liu C.-G., Yang J.-M. Role of MicroRNA miR-27a and miR-451 in the regulation of MDR1/P-glycoprotein expression in human cancer cells. Biochem. Pharmacol. 2008;76:582–588. doi: 10.1016/j.bcp.2008.06.007. PubMed DOI PMC
Vanas V., Haigl B., Stockhammer V., Sutterlüty-Fall H. MicroRNA-21 Increases Proliferation and Cisplatin Sensitivity of Osteosarcoma-Derived Cells. PLoS ONE. 2016;11:e0161023. doi: 10.1371/journal.pone.0161023. PubMed DOI PMC
Douple E.B., Richmond R.C. Platinum complexes as radiosensitizers of hypoxic mammalian cells. Br. J. Cancer Suppl. 1978;3:98–102. PubMed PMC
Boeckman H.J., Trego K.S., Turchi J.J. Cisplatin sensitizes cancer cells to ionizing radiation via inhibition of nonhomologous end joining. Mol. Cancer Res. 2005;3:277–285. doi: 10.1158/1541-7786.MCR-04-0032. PubMed DOI PMC
Ishibashi T., Lippard S.J. Telomere loss in cells treated with cisplatin. Proc. Natl. Acad. Sci. USA. 1998;95:4219–4223. doi: 10.1073/pnas.95.8.4219. PubMed DOI PMC
Jordan P., Carmo-Fonseca M. Cisplatin inhibits synthesis of ribosomal RNA in vivo. Nucleic Acids Res. 1998;26:2831–2836. doi: 10.1093/nar/26.12.2831. PubMed DOI PMC
Tofilon P.J., Vines C.M., Baker F.L., Deen D.F., Brock W.A. cis-Diamminedichloroplatinum(II)-induced sister chromatid exchange: An indicator of sensitivity and heterogeneity in primary human tumor cell cultures. Cancer Res. 1986;46:6156–6159. PubMed
Berndtsson M., Hägg M., Panaretakis T., Havelka A.M., Shoshan M.C., Linder S. Acute apoptosis by cisplatin requires induction of reactive oxygen species but is not associated with damage to nuclear DNA. Int. J. Cancer. 2007;120:175–180. doi: 10.1002/ijc.22132. PubMed DOI
Rouette A., Parent S., Girouard J., Leblanc V., Asselin E. Cisplatin increases B-cell-lymphoma-2 expression via activation of protein kinase C and Akt2 in endometrial cancer cells. Int. J. Cancer. 2012;130:1755–1767. doi: 10.1002/ijc.26183. PubMed DOI
Damia G., Filiberti L., Vikhanskaya F., Carrassa L., Taya Y., D’incalci M., Broggini M. Cisplatinum and taxol induce different patterns of p53 phosphorylation. Neoplasia. 2001;3:10–16. doi: 10.1038/sj.neo.7900122. PubMed DOI PMC
Lützkendorf J., Wieduwild E., Nerger K., Lambrecht N., Schmoll H.-J., Müller-Tidow C., Müller L.P. Resistance for Genotoxic Damage in Mesenchymal Stromal Cells Is Increased by Hypoxia but Not Generally Dependent on p53-Regulated Cell Cycle Arrest. PLoS ONE. 2017;12:e0169921. doi: 10.1371/journal.pone.0169921. PubMed DOI PMC
Sorenson C.M., Barry M.A., Eastman A. Analysis of events associated with cell cycle arrest at G2 phase and cell death induced by cisplatin. J. Natl. Cancer Inst. 1990;82:749–755. doi: 10.1093/jnci/82.9.749. PubMed DOI
Podratz J.L., Knight A.M., Ta L.E., Staff N.P., Gass J.M., Genelin K., Schlattau A., Lathroum L., Windebank A.J. Cisplatin induced mitochondrial DNA damage in dorsal root ganglion neurons. Neurobiol. Dis. 2011;41:661–668. doi: 10.1016/j.nbd.2010.11.017. PubMed DOI PMC
Wagner J.M., Karnitz L.M. Cisplatin-induced DNA damage activates replication checkpoint signaling components that differentially affect tumor cell survival. Mol. Pharmacol. 2009;76:208–214. doi: 10.1124/mol.109.055178. PubMed DOI PMC
Bürkle A., Chen G., Küpper J.H., Grube K., Zeller W.J. Increased poly(ADP-ribosyl)ation in intact cells by cisplatin treatment. Carcinogenesis. 1993;14:559–561. doi: 10.1093/carcin/14.4.559. PubMed DOI
Jennerwein M., Andrews P.A. Drug accumulation and DNA platination in cells exposed to aquated cisplatin species. Cancer Lett. 1994;81:215–220. doi: 10.1016/0304-3835(94)90205-4. PubMed DOI
Shirazi F.H., Molepo J.M., Stewart D.J., Ng C.E., Raaphorst G.P., Goel R. Cytotoxicity, accumulation, and efflux of cisplatin and its metabolites in human ovarian carcinoma cells. Toxicol. Appl. Pharmacol. 1996;140:211–218. doi: 10.1006/taap.1996.0215. PubMed DOI
Hall M.D., Telma K.A., Chang K.-E., Lee T.D., Madigan J.P., Lloyd J.R., Goldlust I.S., Hoeschele J.D., Gottesman M.M. Say no to DMSO: Dimethylsulfoxide inactivates cisplatin, carboplatin, and other platinum complexes. Cancer Res. 2014;74:3913–3922. doi: 10.1158/0008-5472.CAN-14-0247. PubMed DOI PMC
[(accessed on 23 November 2016)]. Available online: http://www.webcitation.org/6mEemW8DM.
Massart C., Le Tellier C., Gibassier J., Leclech G., Nicol M. Modulation by dimethyl sulphoxide of the toxicity induced by cis-diamminedichloroplatinum in cultured thyrocytes. Toxicol. Vitro. 1993;7:87–94. doi: 10.1016/0887-2333(93)90116-M. PubMed DOI
Wiltshaw E., Subramarian S., Alexopoulos C., Barker G.H. Cancer of the ovary: A summary of experience with cis-dichlorodiammineplatinum(II) at the Royal Marsden Hospital. Cancer Treat. Rep. 1979;63:1545–1548. PubMed
Galanski M. Recent developments in the field of anticancer platinum complexes. Recent Pat. Anticancer Drug Discov. 2006;1:285–295. doi: 10.2174/157489206777442287. PubMed DOI
Lebwohl D., Canetta R. Clinical development of platinum complexes in cancer therapy: An historical perspective and an update. Eur. J. Cancer. 1998;34:1522–1534. doi: 10.1016/S0959-8049(98)00224-X. PubMed DOI
Baetz T., Belch A., Couban S., Imrie K., Yau J., Myers R., Ding K., Paul N., Shepherd L., Iglesias J., et al. Gemcitabine, dexamethasone and cisplatin is an active and non-toxic chemotherapy regimen in relapsed or refractory Hodgkin’s disease: A phase II study by the National Cancer Institute of Canada Clinical Trials Group. Ann. Oncol. 2003;14:1762–1767. doi: 10.1093/annonc/mdg496. PubMed DOI
Crump M., Baetz T., Couban S., Belch A., Marcellus D., Howson-Jan K., Imrie K., Myers R., Adams G., Ding K., et al. Gemcitabine, dexamethasone, and cisplatin in patients with recurrent or refractory aggressive histology B-cell non-Hodgkin lymphoma: A Phase II study by the National Cancer Institute of Canada Clinical Trials Group (NCIC-CTG) Cancer. 2004;101:1835–1842. doi: 10.1002/cncr.20587. PubMed DOI
Pearson A.D.J., Pinkerton C.R., Lewis I.J., Imeson J., Ellershaw C., Machin D., European Neuroblastoma Study Group. Children’s Cancer and Leukaemia Group (CCLG formerly United Kingdom Children’s Cancer Study Group) High-dose rapid and standard induction chemotherapy for patients aged over 1 year with stage 4 neuroblastoma: A randomised trial. Lancet Oncol. 2008;9:247–256. PubMed
Reichardt P. The treatment of uterine sarcomas. Ann. Oncol. 2012;23(Suppl. S10):x151–x157. doi: 10.1093/annonc/mds359. PubMed DOI
Dadacaridou M., Papanicolaou X., Maltesas D., Megalakaki C., Patos P., Panteli K., Repousis P., Mitsouli-Mentzikof C. Dexamethasone, cyclophosphamide, etoposide and cisplatin (DCEP) for relapsed or refractory multiple myeloma patients. J. BUON. 2007;12:41–44. PubMed
Glover D., Glick J.H., Weiler C., Fox K., Guerry D. WR-2721 and high-dose cisplatin: An active combination in the treatment of metastatic melanoma. J. Clin. Oncol. 1987;5:574–578. PubMed
Berghmans T., Paesmans M., Lalami Y., Louviaux I., Luce S., Mascaux C., Meert A.P., Sculier J.P. Activity of chemotherapy and immunotherapy on malignant mesothelioma: A systematic review of the literature with meta-analysis. Lung Cancer. 2002;38:111–121. doi: 10.1016/S0169-5002(02)00180-0. PubMed DOI
Hanada K., Nishijima K., Ogata H., Atagi S., Kawahara M. Population pharmacokinetic analysis of cisplatin and its metabolites in cancer patients: Possible misinterpretation of covariates for pharmacokinetic parameters calculated from the concentrations of unchanged cisplatin, ultrafiltered platinum and total platinum. Jpn. J. Clin. Oncol. 2001;31:179–184. PubMed
Daugaard G., Abildgaard U. Cisplatin nephrotoxicity. A review. Cancer Chemother. Pharmacol. 1989;25:1–9. doi: 10.1007/BF00694330. PubMed DOI
Nagai N., Kinoshita M., Ogata H., Tsujino D., Wada Y., Someya K., Ohno T., Masuhara K., Tanaka Y., Kato K., et al. Relationship between pharmacokinetics of unchanged cisplatin and nephrotoxicity after intravenous infusions of cisplatin to cancer patients. Cancer Chemother. Pharmacol. 1996;39:131–137. doi: 10.1007/s002800050548. PubMed DOI
Kartalou M., Essigmann J.M. Mechanisms of resistance to cisplatin. Mutat. Res. 2001;478:23–43. doi: 10.1016/S0027-5107(01)00141-5. PubMed DOI
Olszewski U., Hamilton G. A better platinum-based anticancer drug yet to come? Anticancer Agents Med. Chem. 2010;10:293–301. doi: 10.2174/187152010791162306. PubMed DOI
Dieras V., Girre V., Guilhaume M.-N., Laurence V., Mignot L. Oxaliplatin and ovarian cancer. Bull. Cancer. 2006;93(Suppl. S1):S35–S39. PubMed
Ganjavi H., Gee M., Narendran A., Parkinson N., Krishnamoorthy M., Freedman M.H., Malkin D. Adenovirus-mediated p53 gene therapy in osteosarcoma cell lines: Sensitization to cisplatin and doxorubicin. Cancer Gene Ther. 2006;13:415–419. doi: 10.1038/sj.cgt.7700909. PubMed DOI
Michels J., Vitale I., Senovilla L., Enot D.P., Garcia P., Lissa D., Olaussen K.A., Brenner C., Soria J.-C., Castedo M., et al. Synergistic interaction between cisplatin and PARP inhibitors in non-small cell lung cancer. Cell Cycle. 2013;12:877–883. doi: 10.4161/cc.24034. PubMed DOI PMC
Balmaña J., Tung N.M., Isakoff S.J., Graña B., Ryan P.D., Saura C., Lowe E.S., Frewer P., Winer E., Baselga J., et al. Phase I trial of olaparib in combination with cisplatin for the treatment of patients with advanced breast, ovarian and other solid tumors. Ann. Oncol. 2014;25:1656–1663. doi: 10.1093/annonc/mdu187. PubMed DOI
Sorenson C.M., Eastman A. Influence of cis-diamminedichloroplatinum(II) on DNA synthesis and cell cycle progression in excision repair proficient and deficient Chinese hamster ovary cells. Cancer Res. 1988;48:6703–6707. PubMed
Vichi P., Coin F., Renaud J.P., Vermeulen W., Hoeijmakers J.H., Moras D., Egly J.M. Cisplatin- and UV-damaged DNA lure the basal transcription factor TFIID/TBP. EMBO J. 1997;16:7444–7456. doi: 10.1093/emboj/16.24.7444. PubMed DOI PMC
Cullinane C., Mazur S.J., Essigmann J.M., Phillips D.R., Bohr V.A. Inhibition of RNA polymerase II transcription in human cell extracts by cisplatin DNA damage. Biochemistry. 1999;38:6204–6212. doi: 10.1021/bi982685+. PubMed DOI
Kumar S., Kumar A., Shah P.P., Rai S.N., Panguluri S.K., Kakar S.S. MicroRNA signature of cis-platin resistant vs. cis-platin sensitive ovarian cancer cell lines. J. Ovarian Res. 2011;4:17. PubMed PMC
Ciarimboli G., Ludwig T., Lang D., Pavenstädt H., Koepsell H., Piechota H.-J., Haier J., Jaehde U., Zisowsky J., Schlatter E. Cisplatin nephrotoxicity is critically mediated via the human organic cation transporter 2. Am. J. Pathol. 2005;167:1477–1484. doi: 10.1016/S0002-9440(10)61234-5. PubMed DOI PMC
Gressette M., Vérillaud B., Jimenez-Pailhès A.-S., Lelièvre H., Lo K.-W., Ferrand F.-R., Gattolliat C.-H., Jacquet-Bescond A., Kraus-Berthier L., Depil S., et al. Treatment of Nasopharyngeal Carcinoma Cells with the Histone-Deacetylase Inhibitor Abexinostat: Cooperative Effects with Cis-platin and Radiotherapy on Patient-Derived Xenografts. PLoS ONE. 2014;9:e91325. doi: 10.1371/journal.pone.0091325. PubMed DOI PMC
Rout S.R., Behera B., Maiti T.K., Mohapatra S. Multifunctional magnetic calcium phosphate nanoparticles for targeted platin delivery. Dalton Trans. 2012;41:10777–10783. doi: 10.1039/c2dt30984j. PubMed DOI
Kitao H., Takata M. Fanconi anemia: A disorder defective in the DNA damage response. Int. J. Hematol. 2011;93:417–424. doi: 10.1007/s12185-011-0777-z. PubMed DOI
Sawant A., Kothandapani A., Zhitkovich A., Sobol R.W., Patrick S.M. Role of mismatch repair proteins in the processing of cisplatin interstrand cross-links. DNA Repair. 2015;35:126–136. doi: 10.1016/j.dnarep.2015.10.003. PubMed DOI PMC
Cheng C.H., Kuchta R.D. DNA polymerase epsilon: Aphidicolin inhibition and the relationship between polymerase and exonuclease activity. Biochemistry. 1993;32:8568–8574. doi: 10.1021/bi00084a025. PubMed DOI
Pedrali-Noy G., Spadari S., Miller-Faurès A., Miller A.O., Kruppa J., Koch G. Synchronization of HeLa cell cultures by inhibition of DNA polymerase alpha with aphidicolin. Nucleic Acids Res. 1980;8:377–387. doi: 10.1093/nar/8.2.377. PubMed DOI PMC
Baranovskiy A.G., Babayeva N.D., Suwa Y., Gu J., Pavlov Y.I., Tahirov T.H. Structural basis for inhibition of DNA replication by aphidicolin. Nucleic Acids Res. 2014;42:14013–14021. doi: 10.1093/nar/gku1209. PubMed DOI PMC
Spadari S., Pedrali-Noy G., Falaschi M.C., Ciarrocchi G. Control of DNA replication and cell proliferation in eukaryotes by aphidicolin. Toxicol. Pathol. 1984;12:143–148. doi: 10.1177/019262338401200205. PubMed DOI
[(accessed on 23 January 2017)]. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=4Q5V)
Chang D.J., Lupardus P.J., Cimprich K.A. Monoubiquitination of proliferating cell nuclear antigen induced by stalled replication requires uncoupling of DNA polymerase and mini-chromosome maintenance helicase activities. J. Biol. Chem. 2006;281:32081–32088. doi: 10.1074/jbc.M606799200. PubMed DOI
Sutherland G.R. Chromosomal fragile sites. Genet. Anal. Tech. Appl. 1991;8:161–166. doi: 10.1016/1050-3862(91)90056-W. PubMed DOI
Shiraishi T., Druck T., Mimori K., Flomenberg J., Berk L., Alder H., Miller W., Huebner K., Croce C.M. Sequence conservation at human and mouse orthologous common fragile regions, FRA3B/FHIT and Fra14A2/Fhit. Proc. Natl. Acad. Sci. USA. 2001;98:5722–5727. doi: 10.1073/pnas.091095898. PubMed DOI PMC
Hellman A., Zlotorynski E., Scherer S.W., Cheung J., Vincent J.B., Smith D.I., Trakhtenbrot L., Kerem B. A role for common fragile site induction in amplification of human oncogenes. Cancer Cell. 2002;1:89–97. doi: 10.1016/S1535-6108(02)00017-X. PubMed DOI
Durkin S.G., Ragland R.L., Arlt M.F., Mulle J.G., Warren S.T., Glover T.W. Replication stress induces tumor-like microdeletions in FHIT/FRA3B. Proc. Natl. Acad. Sci. USA. 2008;105:246–251. doi: 10.1073/pnas.0708097105. PubMed DOI PMC
Bristow R.G., Hill R.P. Hypoxia and metabolism. Hypoxia, DNA repair and genetic instability. Nat. Rev. Cancer. 2008;8:180–192. doi: 10.1038/nrc2344. PubMed DOI
MacGregor J.T., Schlegel R., Wehr C.M., Alperin P., Ames B.N. Cytogenetic damage induced by folate deficiency in mice is enhanced by caffeine. Proc. Natl. Acad. Sci. USA. 1990;87:9962–9965. doi: 10.1073/pnas.87.24.9962. PubMed DOI PMC
Koundrioukoff S., Carignon S., Técher H., Letessier A., Brison O., Debatisse M. Stepwise activation of the ATR signaling pathway upon increasing replication stress impacts fragile site integrity. PLoS Genet. 2013;9:e1003643. doi: 10.1371/journal.pgen.1003643. PubMed DOI PMC
Helmrich A., Ballarino M., Tora L. Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol. Cell. 2011;44:966–977. doi: 10.1016/j.molcel.2011.10.013. PubMed DOI
Di Micco R., Fumagalli M., Cicalese A., Piccinin S., Gasparini P., Luise C., Schurra C., Garre’ M., Nuciforo P.G., Bensimon A., et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature. 2006;444:638–642. doi: 10.1038/nature05327. PubMed DOI
Cha R.S., Kleckner N. ATR homolog Mec1 promotes fork progression, thus averting breaks in replication slow zones. Science. 2002;297:602–606. doi: 10.1126/science.1071398. PubMed DOI
Arlt M.F., Mulle J.G., Schaibley V.M., Ragland R.L., Durkin S.G., Warren S.T., Glover T.W. Replication stress induces genome-wide copy number changes in human cells that resemble polymorphic and pathogenic variants. Am. J. Hum. Genet. 2009;84:339–350. doi: 10.1016/j.ajhg.2009.01.024. PubMed DOI PMC
Hardt N., Pedrali-Noy G., Focher F., Spadari S. Aphidicolin does not inhibit DNA repair synthesis in ultraviolet-irradiated HeLa cells. A radioautographic study. Biochem. J. 1981;199:453–455. doi: 10.1042/bj1990453. PubMed DOI PMC
Pedrali-Noy G., Belvedere M., Crepaldi T., Focher F., Spadari S. Inhibition of DNA replication and growth of several human and murine neoplastic cells by aphidicolin without detectable effect upon synthesis of immunoglobulins and HLA antigens. Cancer Res. 1982;42:3810–3813. PubMed
Gera J.F., Fady C., Gardner A., Jacoby F.J., Briskin K.B., Lichtenstein A. Inhibition of DNA repair with aphidicolin enhances sensitivity of targets to tumor necrosis factor. J. Immunol. 1993;151:3746–3757. PubMed
Waters R. Aphidicolin: An inhibitor of DNA repair in human fibroblasts. Carcinogenesis. 1981;2:795–797. doi: 10.1093/carcin/2.8.795. PubMed DOI
Wang F., Stewart J., Price C.M. Human CST abundance determines recovery from diverse forms of DNA damage and replication stress. Cell Cycle. 2014;13:3488–3498. doi: 10.4161/15384101.2014.964100. PubMed DOI PMC
Yeo J.E., Lee E.H., Hendrickson E.A., Sobeck A. CtIP mediates replication fork recovery in a FANCD2-regulated manner. Hum. Mol. Genet. 2014;23:3695–3705. doi: 10.1093/hmg/ddu078. PubMed DOI PMC
Chaudhury I., Stroik D.R., Sobeck A. FANCD2-controlled chromatin access of the Fanconi-associated nuclease FAN1 is crucial for the recovery of stalled replication forks. Mol. Cell. Biol. 2014;34:3939–3954. doi: 10.1128/MCB.00457-14. PubMed DOI PMC
Hammond E.M., Green S.L., Giaccia A.J. Comparison of hypoxia-induced replication arrest with hydroxyurea and aphidicolin-induced arrest. Mutat. Res. 2003;532:205–213. doi: 10.1016/j.mrfmmm.2003.08.017. PubMed DOI
Borel F., Lacroix F.B., Margolis R.L. Prolonged arrest of mammalian cells at the G1/S boundary results in permanent S phase stasis. J. Cell Sci. 2002;115:2829–2838. PubMed
Basile G., Leuzzi G., Pichierri P., Franchitto A. Checkpoint-dependent and independent roles of the Werner syndrome protein in preserving genome integrity in response to mild replication stress. Nucleic Acids Res. 2014;42:12628–12639. doi: 10.1093/nar/gku1022. PubMed DOI PMC
Nguyen G.H., Dexheimer T.S., Rosenthal A.S., Chu W.K., Singh D.K., Mosedale G., Bachrati C.Z., Schultz L., Sakurai M., Savitsky P., et al. A small molecule inhibitor of the BLM helicase modulates chromosome stability in human cells. Chem. Biol. 2013;20:55–62. doi: 10.1016/j.chembiol.2012.10.016. PubMed DOI PMC
Schmidt L., Wiedner M., Velimezi G., Prochazkova J., Owusu M., Bauer S., Loizou J.I. ATMIN is required for the ATM-mediated signaling and recruitment of 53BP1 to DNA damage sites upon replication stress. DNA Repair. 2014;24:122–130. doi: 10.1016/j.dnarep.2014.09.001. PubMed DOI PMC
Fujita M., Sasanuma H., Yamamoto K.N., Harada H., Kurosawa A., Adachi N., Omura M., Hiraoka M., Takeda S., Hirota K. Interference in DNA replication can cause mitotic chromosomal breakage unassociated with double-strand breaks. PLoS ONE. 2013;8:e60043. doi: 10.1371/journal.pone.0060043. PubMed DOI PMC
Beresova L., Vesela E., Chamrad I., Voller J., Yamada M., Furst T., Lenobel R., Chroma K., Gursky J., Krizova K., et al. Role of DNA Repair Factor Xeroderma Pigmentosum Protein Group C in Response to Replication Stress As Revealed by DNA Fragile Site Affinity Chromatography and Quantitative Proteomics. J. Proteome Res. 2016;15:4505–4517. doi: 10.1021/acs.jproteome.6b00622. PubMed DOI
Janson C., Nyhan K., Murnane J.P. Replication Stress and Telomere Dysfunction Are Present in Cultured Human Embryonic Stem Cells. Cytogenet. Genome Res. 2015;146:251–260. doi: 10.1159/000441245. PubMed DOI
Miron K., Golan-Lev T., Dvir R., Ben-David E., Kerem B. Oncogenes create a unique landscape of fragile sites. Nat. Commun. 2015;6:7094. doi: 10.1038/ncomms8094. PubMed DOI
Murfuni I., De Santis A., Federico M., Bignami M., Pichierri P., Franchitto A. Perturbed replication induced genome wide or at common fragile sites is differently managed in the absence of WRN. Carcinogenesis. 2012;33:1655–1663. doi: 10.1093/carcin/bgs206. PubMed DOI
Wilhelm T., Magdalou I., Barascu A., Técher H., Debatisse M., Lopez B.S. Spontaneous slow replication fork progression elicits mitosis alterations in homologous recombination-deficient mammalian cells. Proc. Natl. Acad. Sci. USA. 2014;111:763–768. doi: 10.1073/pnas.1311520111. PubMed DOI PMC
[(accessed on 23 November 2016)]. Available online: https://www.google.cz/url?sa=t&rct=j&q=&esrc=s&source=web&cd=3&cad=rja&uact=8&ved=0ahUKEwibvoX5jL_QAhULVSwKHQfLCXwQFggsMAI&url=https%3A%2F%2Fwww.sigmaaldrich.com%2Fcontent%2Fdam%2Fsigma-aldrich%2Fdocs%2FSigma%2FDatasheet%2F6%2Fa0781dat.pdf&usg=AFQjCNEPSqAi.
Sessa C., Zucchetti M., Davoli E., Califano R., Cavalli F., Frustaci S., Gumbrell L., Sulkes A., Winograd B., D’Incalci M. Phase I and clinical pharmacological evaluation of aphidicolin glycinate. J. Natl. Cancer Inst. 1991;83:1160–1164. doi: 10.1093/jnci/83.16.1160. PubMed DOI
Edelson R.E., Gorycki P.D., MacDonald T.L. The mechanism of aphidicolin bioinactivation by rat liver in vitro systems. Xenobiotica. 1990;20:273–287. doi: 10.3109/00498259009046847. PubMed DOI
Santos G.B., Krogh R., Magalhaes L.G., Andricopulo A.D., Pupo M.T., Emery F.S. Semisynthesis of new aphidicolin derivatives with high activity against Trypanosoma cruzi. Bioorg. Med. Chem. Lett. 2016;26:1205–1208. doi: 10.1016/j.bmcl.2016.01.033. PubMed DOI
Glover T.W., Berger C., Coyle J., Echo B. DNA polymerase alpha inhibition by aphidicolin induces gaps and breaks at common fragile sites in human chromosomes. Hum. Genet. 1984;67:136–142. doi: 10.1007/BF00272988. PubMed DOI
Kurose A., Tanaka T., Huang X., Traganos F., Darzynkiewicz Z. Synchronization in the cell cycle by inhibitors of DNA replication induces histone H2AX phosphorylation: An indication of DNA damage. Cell Prolif. 2006;39:231–240. doi: 10.1111/j.1365-2184.2006.00380.x. PubMed DOI PMC
Trenz K., Smith E., Smith S., Costanzo V. ATM and ATR promote Mre11 dependent restart of collapsed replication forks and prevent accumulation of DNA breaks. EMBO J. 2006;25:1764–1774. doi: 10.1038/sj.emboj.7601045. PubMed DOI PMC
Krakoff I.H., Brown N.C., Reichard P. Inhibition of ribonucleoside diphosphate reductase by hydroxyurea. Cancer Res. 1968;28:1559–1565. PubMed
Reichard P. Interactions between deoxyribonucleotide and DNA synthesis. Annu. Rev. Biochem. 1988;57:349–374. doi: 10.1146/annurev.bi.57.070188.002025. PubMed DOI
Håkansson P., Hofer A., Thelander L. Regulation of mammalian ribonucleotide reduction and dNTP pools after DNA damage and in resting cells. J. Biol. Chem. 2006;281:7834–7841. doi: 10.1074/jbc.M512894200. PubMed DOI
Eriksson M., Uhlin U., Ramaswamy S., Ekberg M., Regnström K., Sjöberg B.M., Eklund H. Binding of allosteric effectors to ribonucleotide reductase protein R1: Reduction of active-site cysteines promotes substrate binding. Structure. 1997;5:1077–1092. doi: 10.1016/S0969-2126(97)00259-1. PubMed DOI
Bianchi V., Pontis E., Reichard P. Changes of deoxyribonucleoside triphosphate pools induced by hydroxyurea and their relation to DNA synthesis. J. Biol. Chem. 1986;261:16037–16042. PubMed
Skog S., Tribukait B., Wallström B., Eriksson S. Hydroxyurea-induced cell death as related to cell cycle in mouse and human T-lymphoma cells. Cancer Res. 1987;47:6490–6493. PubMed
Akerblom L. Azidocytidine is incorporated into RNA of 3T6 mouse fibroblasts. FEBS Lett. 1985;193:203–207. doi: 10.1016/0014-5793(85)80151-4. PubMed DOI
Anglana M., Apiou F., Bensimon A., Debatisse M. Dynamics of DNA replication in mammalian somatic cells: Nucleotide pool modulates origin choice and interorigin spacing. Cell. 2003;114:385–394. doi: 10.1016/S0092-8674(03)00569-5. PubMed DOI
Barlow J.H., Faryabi R.B., Callén E., Wong N., Malhowski A., Chen H.T., Gutierrez-Cruz G., Sun H.-W., McKinnon P., Wright G., et al. Identification of early replicating fragile sites that contribute to genome instability. Cell. 2013;152:620–632. doi: 10.1016/j.cell.2013.01.006. PubMed DOI PMC
Lönn U., Lönn S. Extensive regions of single-stranded DNA in aphidicolin-treated melanoma cells. Biochemistry. 1988;27:566–570. doi: 10.1021/bi00402a010. PubMed DOI
Recolin B., Van der Laan S., Maiorano D. Role of replication protein A as sensor in activation of the S-phase checkpoint in Xenopus egg extracts. Nucleic Acids Res. 2012;40:3431–3442. doi: 10.1093/nar/gkr1241. PubMed DOI PMC
Arlt M.F., Ozdemir A.C., Birkeland S.R., Wilson T.E., Glover T.W. Hydroxyurea induces de novo copy number variants in human cells. Proc. Natl. Acad. Sci. USA. 2011;108:17360–17365. doi: 10.1073/pnas.1109272108. PubMed DOI PMC
Huang M.-E., Facca C., Fatmi Z., Baïlle D., Bénakli S., Vernis L. DNA replication inhibitor hydroxyurea alters Fe–S centers by producing reactive oxygen species in vivo. Sci. Rep. 2016;6:29361. doi: 10.1038/srep29361. PubMed DOI PMC
Szikriszt B., Póti Á., Pipek O., Krzystanek M., Kanu N., Molnár J., Ribli D., Szeltner Z., Tusnády G.E., Csabai I., et al. A comprehensive survey of the mutagenic impact of common cancer cytotoxics. Genome Biol. 2016;17:99. doi: 10.1186/s13059-016-0963-7. PubMed DOI PMC
Mistrik M., Oplustilova L., Lukas J., Bartek J. Low-dose DNA damage and replication stress responses quantified by optimized automated single-cell image analysis. Cell Cycle. 2009;8:2592–2599. doi: 10.4161/cc.8.16.9331. PubMed DOI
Ohouo P.Y., Bastos de Oliveira F.M., Liu Y., Ma C.J., Smolka M.B. DNA-repair scaffolds dampen checkpoint signalling by counteracting the adaptor Rad9. Nature. 2013;493:120–124. doi: 10.1038/nature11658. PubMed DOI PMC
Morafraile E.C., Diffley J.F.X., Tercero J.A., Segurado M. Checkpoint-dependent RNR induction promotes fork restart after replicative stress. Sci. Rep. 2015;5:7886. doi: 10.1038/srep07886. PubMed DOI PMC
Kim H.-S., Kim S.-K., Hromas R., Lee S.-H. The SET Domain Is Essential for Metnase Functions in Replication Restart and the 5’ End of SS-Overhang Cleavage. PLoS ONE. 2015;10:e0139418. doi: 10.1371/journal.pone.0139418. PubMed DOI PMC
Masuda T., Xu X., Dimitriadis E.K., Lahusen T., Deng C.-X. “DNA Binding Region” of BRCA1 Affects Genetic Stability through modulating the Intra-S-Phase Checkpoint. Int. J. Biol. Sci. 2016;12:133–143. doi: 10.7150/ijbs.14242. PubMed DOI PMC
Yarden R.I., Metsuyanim S., Pickholtz I., Shabbeer S., Tellio H., Papa M.Z. BRCA1-dependent Chk1 phosphorylation triggers partial chromatin disassociation of phosphorylated Chk1 and facilitates S-phase cell cycle arrest. Int. J. Biochem. Cell Biol. 2012;44:1761–1769. doi: 10.1016/j.biocel.2012.06.026. PubMed DOI PMC
Awate S., De Benedetti A. TLK1B mediated phosphorylation of Rad9 regulates its nuclear/cytoplasmic localization and cell cycle checkpoint. BMC Mol. Biol. 2016;17:3. doi: 10.1186/s12867-016-0056-x. PubMed DOI PMC
Ahlskog J.K., Larsen B.D., Achanta K., Sørensen C.S. ATM/ATR-mediated phosphorylation of PALB2 promotes RAD51 function. EMBO Rep. 2016;17:671–681. doi: 10.15252/embr.201541455. PubMed DOI PMC
Molina B., Marchetti F., Gómez L., Ramos S., Torres L., Ortiz R., Altamirano-Lozano M., Carnevale A., Frias S. Hydroxyurea induces chromosomal damage in G2 and enhances the clastogenic effect of mitomycin C in Fanconi anemia cells. Environ. Mol. Mutagen. 2015;56:457–467. doi: 10.1002/em.21938. PubMed DOI
Croke M., Neumann M.A., Grotsky D.A., Kreienkamp R., Yaddanapudi S.C., Gonzalo S. Differences in 53BP1 and BRCA1 regulation between cycling and non-cycling cells. Cell Cycle. 2013;12:3629–3639. doi: 10.4161/cc.26582. PubMed DOI PMC
Yamada M., Watanabe K., Mistrik M., Vesela E., Protivankova I., Mailand N., Lee M., Masai H., Lukas J., Bartek J. ATR-Chk1-APC/CCdh1-dependent stabilization of Cdc7-ASK (Dbf4) kinase is required for DNA lesion bypass under replication stress. Genes Dev. 2013;27:2459–2472. doi: 10.1101/gad.224568.113. PubMed DOI PMC
Hu L., Kim T.M., Son M.Y., Kim S.-A., Holland C.L., Tateishi S., Kim D.H., Yew P.R., Montagna C., Dumitrache L.C., et al. Two replication fork maintenance pathways fuse inverted repeats to rearrange chromosomes. Nature. 2013;501:569–572. doi: 10.1038/nature12500. PubMed DOI PMC
Lou T.-F., Singh M., Mackie A., Li W., Pace B.S. Hydroxyurea generates nitric oxide in human erythroid cells: Mechanisms for gamma-globin gene activation. Exp. Biol. Med. 2009;234:1374–1382. doi: 10.3181/0811-RM-339. PubMed DOI PMC
Vassileva I., Yanakieva I., Peycheva M., Gospodinov A., Anachkova B. The mammalian INO80 chromatin remodeling complex is required for replication stress recovery. Nucleic Acids Res. 2014;42:9074–9086. doi: 10.1093/nar/gku605. PubMed DOI PMC
Park J.I., Choi H.S., Jeong J.S., Han J.Y., Kim I.H. Involvement of p38 kinase in hydroxyurea-induced differentiation of K562 cells. Cell Growth Differ. 2001;12:481–486. PubMed
Barthelemy J., Hanenberg H., Leffak M. FANCJ is essential to maintain microsatellite structure genome-wide during replication stress. Nucleic Acids Res. 2016;44:6803–6816. doi: 10.1093/nar/gkw433. PubMed DOI PMC
Kunnev D., Rusiniak M.E., Kudla A., Freeland A., Cady G.K., Pruitt S.C. DNA damage response and tumorigenesis in Mcm2-deficient mice. Oncogene. 2010;29:3630–3638. doi: 10.1038/onc.2010.125. PubMed DOI PMC
Da Guarda C.C., Santiago R.P., Pitanga T.N., Santana S.S., Zanette D.L., Borges V.M., Goncalves M.S. Heme changes HIF-α, eNOS and nitrite production in HUVECs after simvastatin, HU, and ascorbic acid therapies. Microvasc. Res. 2016;106:128–136. doi: 10.1016/j.mvr.2016.04.002. PubMed DOI
Leitch C., Osdal T., Andresen V., Molland M., Kristiansen S., Nguyen X.N., Bruserud Ø., Gjertsen B.T., McCormack E. Hydroxyurea synergizes with valproic acid in wild-type p53 acute myeloid leukaemia. Oncotarget. 2016;7:8105–8118. PubMed PMC
Liu K., Graves J.D., Scott J.D., Li R., Lin W.-C. Akt switches TopBP1 function from checkpoint activation to transcriptional regulation through phosphoserine binding-mediated oligomerization. Mol. Cell. Biol. 2013;33:4685–4700. doi: 10.1128/MCB.00373-13. PubMed DOI PMC
[(accessed on 23 January 2017)]. Available online: https://www.google.cz/url?sa=t&rct=j&q=&esrc=s&source=web&cd=3&cad=rja&uact=8&ved=0ahUKEwiVt4Lulb_QAhUBGSwKHbcOB_kQFggsMAI&url=https%3A%2F%2Fwww.sigmaaldrich.com%2Fcontent%2Fdam%2Fsigma-aldrich%2Fdocs%2FSigma%2FProduct_Information_Sheet%2F2%2Fh8627pis.pdf&.
Segal J.B., Strouse J.J., Beach M.C., Haywood C., Witkop C., Park H., Wilson R.F., Bass E.B., Lanzkron S. Hydroxyurea for the Treatment of Sickle Cell Disease. Agency for Healthcare Research and Quality (US); Rockville, MD, USA: 2008. pp. 1–95. PubMed PMC
Kühr T., Burgstaller S., Apfelbeck U., Linkesch W., Seewann H., Fridrik M., Michlmayr G., Krieger O., Lutz D., Lin W., et al. A randomized study comparing interferon (IFNα) plus low-dose cytarabine and interferon plus hydroxyurea (HU) in early chronic-phase chronic myeloid leukemia (CML) Leuk. Res. 2003;27:405–411. doi: 10.1016/S0145-2126(02)00223-0. PubMed DOI
Aruch D., Mascarenhas J. Contemporary approach to essential thrombocythemia and polycythemia vera. Curr. Opin. Hematol. 2016;23:150–160. doi: 10.1097/MOH.0000000000000216. PubMed DOI
Barbui T., Finazzi M.C., Finazzi G. Front-line therapy in polycythemia vera and essential thrombocythemia. Blood Rev. 2012;26:205–211. doi: 10.1016/j.blre.2012.06.002. PubMed DOI
Benito J.M., López M., Lozano S., Ballesteros C., González-Lahoz J., Soriano V. Hydroxyurea exerts an anti-proliferative effect on T cells but has no direct impact on cellular activation. Clin. Exp. Immunol. 2007;149:171–177. doi: 10.1111/j.1365-2249.2007.03412.x. PubMed DOI PMC
Gurberg J., Bouganim N., Shenouda G., Zeitouni A. A case of recurrent anaplastic meningioma of the skull base with radiologic response to hydroxyurea. J. Neurol. Surg. Rep. 2014;75:e52–e55. doi: 10.1055/s-0033-1359300. PubMed DOI PMC
Kiladjian J.-J., Chevret S., Dosquet C., Chomienne C., Rain J.-D. Treatment of polycythemia vera with hydroxyurea and pipobroman: Final results of a randomized trial initiated in 1980. J. Clin. Oncol. 2011;29:3907–3913. doi: 10.1200/JCO.2011.36.0792. PubMed DOI
Charache S., Barton F.B., Moore R.D., Terrin M.L., Steinberg M.H., Dover G.J., Ballas S.K., McMahon R.P., Castro O., Orringer E.P. Hydroxyurea and sickle cell anemia. Clinical utility of a myelosuppressive “switching” agent. The Multicenter Study of Hydroxyurea in Sickle Cell Anemia. Medicine. 1996;75:300–326. doi: 10.1097/00005792-199611000-00002. PubMed DOI
Steinberg M.H., McCarthy W.F., Castro O., Ballas S.K., Armstrong F.D., Smith W., Ataga K., Swerdlow P., Kutlar A., DeCastro L., et al. The risks and benefits of long-term use of hydroxyurea in sickle cell anemia: A 17.5 year follow-up. Am. J. Hematol. 2010;85:403–408. doi: 10.1002/ajh.21699. PubMed DOI PMC
Darzynkiewicz Z., Halicka H.D., Zhao H., Podhorecka M. Cell synchronization by inhibitors of DNA replication induces replication stress and DNA damage response: Analysis by flow cytometry. Methods Mol. Biol. 2011;761:85–96. PubMed PMC
Fugger K., Mistrik M., Danielsen J.R., Dinant C., Falck J., Bartek J., Lukas J., Mailand N. Human Fbh1 helicase contributes to genome maintenance via pro- and anti-recombinase activities. J. Cell Biol. 2009;186:655–663. doi: 10.1083/jcb.200812138. PubMed DOI PMC
Liu N., Lim C.-S. Differential roles of XRCC2 in homologous recombinational repair of stalled replication forks. J. Cell. Biochem. 2005;95:942–954. doi: 10.1002/jcb.20457. PubMed DOI
Brose R.D., Shin G., McGuinness M.C., Schneidereith T., Purvis S., Dong G.X., Keefer J., Spencer F., Smith K.D. Activation of the stress proteome as a mechanism for small molecule therapeutics. Hum. Mol. Genet. 2012;21:4237–4252. doi: 10.1093/hmg/dds247. PubMed DOI PMC
Adragna N.C., Fonseca P., Lauf P.K. Hydroxyurea affects cell morphology, cation transport, and red blood cell adhesion in cultured vascular endothelial cells. Blood. 1994;83:553–560. PubMed
Wall M.E., Wani M.C., Cook C.E., Palmer K.H., McPhail A.T., Sim G.A. Plant Antitumor Agents. I. The Isolation and Structure of Camptothecin, a Novel Alkaloidal Leukemia and Tumor Inhibitor from Camptotheca acuminata1,2. J. Am. Chem. Soc. 1966;88:3888–3890. doi: 10.1021/ja00968a057. DOI
Gupta M., Fujimori A., Pommier Y. Eukaryotic DNA topoisomerases I. Biochim. Biophys. Acta. 1995;1262:1–14. doi: 10.1016/0167-4781(95)00029-G. PubMed DOI
Champoux J.J. Mechanism of the reaction catalyzed by the DNA untwisting enzyme: Attachment of the enzyme to 3′-terminus of the nicked DNA. J. Mol. Biol. 1978;118:441–446. doi: 10.1016/0022-2836(78)90238-3. PubMed DOI
[(accessed on 23 January 2017)]. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=1T8I.
Stivers J.T., Harris T.K., Mildvan A.S. Vaccinia DNA topoisomerase I: Evidence supporting a free rotation mechanism for DNA supercoil relaxation. Biochemistry. 1997;36:5212–5222. doi: 10.1021/bi962880t. PubMed DOI
Koster D.A., Palle K., Bot E.S.M., Bjornsti M.-A., Dekker N.H. Antitumour drugs impede DNA uncoiling by topoisomerase I. Nature. 2007;448:213–217. doi: 10.1038/nature05938. PubMed DOI
Staker B.L., Hjerrild K., Feese M.D., Behnke C.A., Burgin A.B., Stewart L. The mechanism of topoisomerase I poisoning by a camptothecin analog. Proc. Natl. Acad. Sci. USA. 2002;99:15387–15392. doi: 10.1073/pnas.242259599. PubMed DOI PMC
Regairaz M., Zhang Y.-W., Fu H., Agama K.K., Tata N., Agrawal S., Aladjem M.I., Pommier Y. Mus81-mediated DNA cleavage resolves replication forks stalled by topoisomerase I–DNA complexes. J. Cell Biol. 2011;195:739–749. doi: 10.1083/jcb.201104003. PubMed DOI PMC
Palle K., Vaziri C. Rad18 E3 ubiquitin ligase activity mediates Fanconi anemia pathway activation and cell survival following DNA Topoisomerase 1 inhibition. Cell Cycle. 2011;10:1625–1638. doi: 10.4161/cc.10.10.15617. PubMed DOI PMC
Pommier Y. Topoisomerase I inhibitors: Camptothecins and beyond. Nat. Rev. Cancer. 2006;6:789–802. doi: 10.1038/nrc1977. PubMed DOI
Tuduri S., Crabbé L., Conti C., Tourrière H., Holtgreve-Grez H., Jauch A., Pantesco V., De Vos J., Thomas A., Theillet C., et al. Topoisomerase I suppresses genomic instability by preventing interference between replication and transcription. Nat. Cell Biol. 2009;11:1315–1324. doi: 10.1038/ncb1984. PubMed DOI PMC
Tripathi K., Mani C., Clark D.W., Palle K. Rad18 is required for functional interactions between FANCD2, BRCA2, and Rad51 to repair DNA topoisomerase 1-poisons induced lesions and promote fork recovery. Oncotarget. 2016;7:12537–12553. PubMed PMC
Tsao Y.P., D’Arpa P., Liu L.F. The involvement of active DNA synthesis in camptothecin-induced G2 arrest: Altered regulation of p34cdc2/cyclin B. Cancer Res. 1992;52:1823–1829. PubMed
Kharbanda S., Rubin E., Gunji H., Hinz H., Giovanella B., Pantazis P., Kufe D. Camptothecin and its derivatives induce expression of the c-jun protooncogene in human myeloid leukemia cells. Cancer Res. 1991;51:6636–6642. PubMed
Aller P., Rius C., Mata F., Zorrilla A., Cabañas C., Bellón T., Bernabeu C. Camptothecin induces differentiation and stimulates the expression of differentiation-related genes in U-937 human promonocytic leukemia cells. Cancer Res. 1992;52:1245–1251. PubMed
Clements M.K., Jones C.B., Cumming M., Daoud S.S. Antiangiogenic potential of camptothecin and topotecan. Cancer Chemother. Pharmacol. 1999;44:411–416. doi: 10.1007/s002800050997. PubMed DOI
O’Leary J.J., Shapiro R.L., Ren C.J., Chuang N., Cohen H.W., Potmesil M. Antiangiogenic effects of camptothecin analogues 9-amino-20(S)-camptothecin, topotecan, and CPT-11 studied in the mouse cornea model. Clin. Cancer Res. 1999;5:181–187. PubMed
Arlt M.F., Glover T.W. Inhibition of topoisomerase I prevents chromosome breakage at common fragile sites. DNA Repair. 2010;9:678–689. doi: 10.1016/j.dnarep.2010.03.005. PubMed DOI PMC
Horwitz S.B., Horwitz M.S. Effects of camptothecin on the breakage and repair of DNA during the cell cycle. Cancer Res. 1973;33:2834–2836. PubMed
Jayasooriya R.G.P.T., Choi Y.H., Hyun J.W., Kim G.-Y. Camptothecin sensitizes human hepatoma Hep3B cells to TRAIL-mediated apoptosis via ROS-dependent death receptor 5 upregulation with the involvement of MAPKs. Environ. Toxicol. Pharmacol. 2014;38:959–967. doi: 10.1016/j.etap.2014.10.012. PubMed DOI
Strumberg D., Pilon A.A., Smith M., Hickey R., Malkas L., Pommier Y. Conversion of topoisomerase I cleavage complexes on the leading strand of ribosomal DNA into 5’-phosphorylated DNA double-strand breaks by replication runoff. Mol. Cell. Biol. 2000;20:3977–3987. doi: 10.1128/MCB.20.11.3977-3987.2000. PubMed DOI PMC
Priel E., Showalter S.D., Roberts M., Oroszlan S., Blair D.G. The topoisomerase I inhibitor, camptothecin, inhibits equine infectious anemia virus replication in chronically infected CF2Th cells. J. Virol. 1991;65:4137–4141. PubMed PMC
Bruno S., Giaretti W., Darzynkiewicz Z. Effect of camptothecin on mitogenic stimulation of human lymphocytes: Involvement of DNA topoisomerase I in cell transition from G0 to G1 phase of the cell cycle and in DNA replication. J. Cell. Physiol. 1992;151:478–486. doi: 10.1002/jcp.1041510306. PubMed DOI
Squires S., Ryan A.J., Strutt H.L., Johnson R.T. Hypersensitivity of Cockayne’s syndrome cells to camptothecin is associated with the generation of abnormally high levels of double strand breaks in nascent DNA. Cancer Res. 1993;53:2012–2019. PubMed
Ding X., Matsuo K., Xu L., Yang J., Zheng L. Optimized combinations of bortezomib, camptothecin, and doxorubicin show increased efficacy and reduced toxicity in treating oral cancer. Anticancer Drugs. 2015;26:547–554. doi: 10.1097/CAD.0000000000000222. PubMed DOI
Zhang J., Walter J.C. Mechanism and regulation of incisions during DNA interstrand cross-link repair. DNA Repair. 2014;19:135–142. doi: 10.1016/j.dnarep.2014.03.018. PubMed DOI PMC
Ray Chaudhuri A., Hashimoto Y., Herrador R., Neelsen K.J., Fachinetti D., Bermejo R., Cocito A., Costanzo V., Lopes M. Topoisomerase I poisoning results in PARP-mediated replication fork reversal. Nat. Struct. Mol. Biol. 2012;19:417–423. doi: 10.1038/nsmb.2258. PubMed DOI
[(accessed on 23 November 2016)]. Available online: http://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Sigma/Product_Information_Sheet/c9911pis.pdf.
Jaxel C., Kohn K.W., Wani M.C., Wall M.E., Pommier Y. Structure-activity study of the actions of camptothecin derivatives on mammalian topoisomerase I: Evidence for a specific receptor site and a relation to antitumor activity. Cancer Res. 1989;49:1465–1469. PubMed
Takagi K., Dexheimer T.S., Redon C., Sordet O., Agama K., Lavielle G., Pierré A., Bates S.E., Pommier Y. Novel E-ring camptothecin keto analogues (S38809 and S39625) are stable, potent, and selective topoisomerase I inhibitors without being substrates of drug efflux transporters. Mol. Cancer Ther. 2007;6:3229–3238. doi: 10.1158/1535-7163.MCT-07-0441. PubMed DOI
Hande K.R. Etoposide: Four decades of development of a topoisomerase II inhibitor. Eur. J. Cancer. 1998;34:1514–1521. doi: 10.1016/S0959-8049(98)00228-7. PubMed DOI
[(accessed on 23 January 2017)]. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=3QX3.
Liu L.F., Rowe T.C., Yang L., Tewey K.M., Chen G.L. Cleavage of DNA by mammalian DNA topoisomerase II. J. Biol. Chem. 1983;258:15365–15370. PubMed
Gibson E.G., King M.M., Mercer S.L., Deweese J.E. Two-Mechanism Model for the Interaction of Etoposide Quinone with Topoisomerase IIα. Chem. Res. Toxicol. 2016;29:1541–1548. doi: 10.1021/acs.chemrestox.6b00209. PubMed DOI
Wu C.-C., Li T.-K., Farh L., Lin L.-Y., Lin T.-S., Yu Y.-J., Yen T.-J., Chiang C.-W., Chan N.-L. Structural basis of type II topoisomerase inhibition by the anticancer drug etoposide. Science. 2011;333:459–462. doi: 10.1126/science.1204117. PubMed DOI
Bender R.P., Jablonksy M.J., Shadid M., Romaine I., Dunlap N., Anklin C., Graves D.E., Osheroff N. Substituents on etoposide that interact with human topoisomerase IIalpha in the binary enzyme-drug complex: Contributions to etoposide binding and activity. Biochemistry. 2008;47:4501–4509. doi: 10.1021/bi702019z. PubMed DOI PMC
Wilstermann A.M., Bender R.P., Godfrey M., Choi S., Anklin C., Berkowitz D.B., Osheroff N., Graves D.E. Topoisomerase II—Drug interaction domains: Identification of substituents on etoposide that interact with the enzyme. Biochemistry. 2007;46:8217–8225. doi: 10.1021/bi700272u. PubMed DOI PMC
Jacob D.A., Mercer S.L., Osheroff N., Deweese J.E. Etoposide quinone is a redox-dependent topoisomerase II poison. Biochemistry. 2011;50:5660–5667. doi: 10.1021/bi200438m. PubMed DOI PMC
Rogakou E.P., Pilch D.R., Orr A.H., Ivanova V.S., Bonner W.M. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J. Biol. Chem. 1998;273:5858–5868. doi: 10.1074/jbc.273.10.5858. PubMed DOI
Terasawa M., Shinohara A., Shinohara M. Canonical non-homologous end joining in mitosis induces genome instability and is suppressed by M-phase-specific phosphorylation of XRCC4. PLoS Genet. 2014;10:e1004563. doi: 10.1371/journal.pgen.1004563. PubMed DOI PMC
Zhao H., Rybak P., Dobrucki J., Traganos F., Darzynkiewicz Z. Relationship of DNA damage signaling to DNA replication following treatment with DNA topoisomerase inhibitors camptothecin/topotecan, mitoxantrone, or etoposide. Cytometry A. 2012;81:45–51. doi: 10.1002/cyto.a.21172. PubMed DOI PMC
Montecucco A., Rossi R., Ferrari G., Scovassi A.I., Prosperi E., Biamonti G. Etoposide Induces the Dispersal of DNA Ligase I from Replication Factories. Mol. Biol. Cell. 2001;12:2109–2118. doi: 10.1091/mbc.12.7.2109. PubMed DOI PMC
Holm C., Covey J.M., Kerrigan D., Pommier Y. Differential requirement of DNA replication for the cytotoxicity of DNA topoisomerase I and II inhibitors in Chinese hamster DC3F cells. Cancer Res. 1989;49:6365–6368. PubMed
Austin C.A., Sng J.H., Patel S., Fisher L.M. Novel HeLa topoisomerase II is the II beta isoform: Complete coding sequence and homology with other type II topoisomerases. Biochim. Biophys. Acta. 1993;1172:283–291. doi: 10.1016/0167-4781(93)90215-Y. PubMed DOI
Niimi A., Suka N., Harata M., Kikuchi A., Mizuno S. Co-localization of chicken DNA topoisomerase IIalpha, but not beta, with sites of DNA replication and possible involvement of a C-terminal region of alpha through its binding to PCNA. Chromosoma. 2001;110:102–114. doi: 10.1007/s004120100140. PubMed DOI
Ju B.-G., Lunyak V.V., Perissi V., Garcia-Bassets I., Rose D.W., Glass C.K., Rosenfeld M.G. A topoisomerase IIβ-mediated dsDNA break required for regulated transcription. Science. 2006;312:1798–1802. doi: 10.1126/science.1127196. PubMed DOI
Azarova A.M., Lyu Y.L., Lin C.-P., Tsai Y.-C., Lau J.Y.-N., Wang J.C., Liu L.F. Roles of DNA topoisomerase II isozymes in chemotherapy and secondary malignancies. Proc. Natl. Acad. Sci. USA. 2007;104:11014–11019. doi: 10.1073/pnas.0704002104. PubMed DOI PMC
Nitiss J.L. DNA topoisomerase II and its growing repertoire of biological functions. Nat. Rev. Cancer. 2009;9:327–337. doi: 10.1038/nrc2608. PubMed DOI PMC
Gupta R.S., Bromke A., Bryant D.W., Gupta R., Singh B., McCalla D.R. Etoposide (VP16) and teniposide (VM26): Novel anticancer drugs, strongly mutagenic in mammalian but not prokaryotic test systems. Mutagenesis. 1987;2:179–186. doi: 10.1093/mutage/2.3.179. PubMed DOI
Muslimović A., Nyström S., Gao Y., Hammarsten O. Numerical Analysis of Etoposide Induced DNA Breaks. PLoS ONE. 2009;4:e5859. doi: 10.1371/annotation/290cebfd-d5dc-4bd2-99b4-f4cf0be6c838. PubMed DOI PMC
Álvarez-Quilón A., Serrano-Benítez A., Lieberman J.A., Quintero C., Sánchez-Gutiérrez D., Escudero L.M., Cortés-Ledesma F. ATM specifically mediates repair of double-strand breaks with blocked DNA ends. Nat. Commun. 2014;5:3347. doi: 10.1038/ncomms4347. PubMed DOI PMC
Nagano T., Nakano M., Nakashima A., Onishi K., Yamao S., Enari M., Kikkawa U., Kamada S. Identification of cellular senescence-specific genes by comparative transcriptomics. Sci. Rep. 2016;6:31758. doi: 10.1038/srep31758. PubMed DOI PMC
Brasacchio D., Alsop A.E., Noori T., Lufti M., Iyer S., Simpson K.J., Bird P.I., Kluck R.M., Johnstone R.W., Trapani J.A. Epigenetic control of mitochondrial cell death through PACS1-mediated regulation of BAX/BAK oligomerization. Cell Death Differ. 2017 doi: 10.1038/cdd.2016.119. PubMed DOI PMC
Martin R., Desponds C., Eren R.O., Quadroni M., Thome M., Fasel N. Caspase-mediated cleavage of raptor participates in the inactivation of mTORC1 during cell death. Cell Death Discov. 2016;2:16024. doi: 10.1038/cddiscovery.2016.24. PubMed DOI PMC
Brekman A., Singh K.E., Polotskaia A., Kundu N., Bargonetti J. A p53-independent role of Mdm2 in estrogen-mediated activation of breast cancer cell proliferation. Breast Cancer Res. 2011;13:R3. doi: 10.1186/bcr2804. PubMed DOI PMC
Soubeyrand S., Pope L., Haché R.J.G. Topoisomerase IIα-dependent induction of a persistent DNA damage response in response to transient etoposide exposure. Mol. Oncol. 2010;4:38–51. doi: 10.1016/j.molonc.2009.09.003. PubMed DOI PMC
Velma V., Carrero Z.I., Allen C.B., Hebert M.D. Coilin levels modulate cell cycle progression and γH2AX levels in etoposide treated U2OS cells. FEBS Lett. 2012;586:3404–3409. doi: 10.1016/j.febslet.2012.07.054. PubMed DOI PMC
Dehennaut V., Loison I., Dubuissez M., Nassour J., Abbadie C., Leprince D. DNA double-strand breaks lead to activation of hypermethylated in cancer 1 (HIC1) by SUMOylation to regulate DNA repair. J. Biol. Chem. 2013;288:10254–10264. doi: 10.1074/jbc.M112.421610. PubMed DOI PMC
Paget S., Dubuissez M., Dehennaut V., Nassour J., Harmon B.T., Spruyt N., Loison I., Abbadie C., Rood B.R., Leprince D. HIC1 (hypermethylated in cancer 1) SUMOylation is dispensable for DNA repair but is essential for the apoptotic DNA damage response (DDR) to irreparable DNA double-strand breaks (DSBs) Oncotarget. 2017;8:2916–2935. doi: 10.18632/oncotarget.13807. PubMed DOI PMC
Sypniewski D., Bednarek I., Gałka S., Loch T., Błaszczyk D., Sołtysik D. Cytotoxicity of etoposide in cancer cell lines in vitro after BCL-2 and C-RAF gene silencing with antisense oligonucleotides. Acta Pol. Pharm. 2013;70:87–97. PubMed
Rybak P., Hoang A., Bujnowicz L., Bernas T., Berniak K., ZarÄ Bski M.A., Darzynkiewicz Z., Dobrucki J. Low level phosphorylation of histone H2AX on serine 139 (γH2AX) is not associated with DNA double-strand breaks. Oncotarget. 2016;7:49574–49587. doi: 10.18632/oncotarget.10411. PubMed DOI PMC
Chen L., Cui H., Fang J., Deng H., Kuang P., Guo H., Wang X., Zhao L. Glutamine deprivation plus BPTES alters etoposide- and cisplatin-induced apoptosis in triple negative breast cancer cells. Oncotarget. 2016;7:54691–54701. doi: 10.18632/oncotarget.10579. PubMed DOI PMC
Rodriguez-Lopez A.M., Xenaki D., Eden T.O., Hickman J.A., Chresta C.M. MDM2 mediated nuclear exclusion of p53 attenuates etoposide-induced apoptosis in neuroblastoma cells. Mol. Pharmacol. 2001;59:135–143. PubMed
Litwiniec A., Gackowska L., Helmin-Basa A., Żuryń A., Grzanka A. Low-dose etoposide-treatment induces endoreplication and cell death accompanied by cytoskeletal alterations in A549 cells: Does the response involve senescence? The possible role of vimentin. Cancer Cell Int. 2013;13:9. doi: 10.1186/1475-2867-13-9. PubMed DOI PMC
Akhtar N., Talegaonkar S., Khar R.K., Jaggi M. A validated stability-indicating LC method for estimation of etoposide in bulk and optimized self-nano emulsifying formulation: Kinetics and stability effects. Saudi Pharm. J. 2013;21:103–111. doi: 10.1016/j.jsps.2012.01.005. PubMed DOI PMC
[(accessed on 23 Novermber 2016)]. Available online: http://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Sigma/Product_Information_Sheet/e1383pis.pdf.
Wrasidlo W., Schröder U., Bernt K., Hübener N., Shabat D., Gaedicke G., Lode H. Synthesis, hydrolytic activation and cytotoxicity of etoposide prodrugs. Bioorg Med Chem Lett. 2002;12:557–560. doi: 10.1016/S0960-894X(01)00801-0. PubMed DOI
Jokić M., Vlašić I., Rinneburger M., Klümper N., Spiro J., Vogel W., Offermann A., Kümpers C., Fritz C., Schmitt A., et al. Ercc1 Deficiency Promotes Tumorigenesis and Increases Cisplatin Sensitivity in a Tp53 Context-Specific Manner. Mol. Cancer Res. 2016;14:1110–1123. doi: 10.1158/1541-7786.MCR-16-0094. PubMed DOI
Felix C.A., Walker A.H., Lange B.J., Williams T.M., Winick N.J., Cheung N.K., Lovett B.D., Nowell P.C., Blair I.A., Rebbeck T.R. Association of CYP3A4 genotype with treatment-related leukemia. Proc. Natl. Acad. Sci. USA. 1998;95:13176–13181. doi: 10.1073/pnas.95.22.13176. PubMed DOI PMC
Blanco J.G., Edick M.J., Relling M.V. Etoposide induces chimeric Mll gene fusions. FASEB J. 2004;18:173–175. doi: 10.1096/fj.03-0638fje. PubMed DOI
Thirman M.J., Gill H.J., Burnett R.C., Mbangkollo D., McCabe N.R., Kobayashi H., Ziemin-van der Poel S., Kaneko Y., Morgan R., Sandberg A.A. Rearrangement of the MLL gene in acute lymphoblastic and acute myeloid leukemias with 11q23 chromosomal translocations. N. Engl. J. Med. 1993;329:909–914. doi: 10.1056/NEJM199309233291302. PubMed DOI
Cerveira N., Lisboa S., Correia C., Bizarro S., Santos J., Torres L., Vieira J., Barros-Silva J.D., Pereira D., Moreira C., et al. Genetic and clinical characterization of 45 acute leukemia patients with MLL gene rearrangements from a single institution. Mol. Oncol. 2012;6:553–564. doi: 10.1016/j.molonc.2012.06.004. PubMed DOI PMC
Krivtsov A.V., Armstrong S.A. MLL translocations, histone modifications and leukaemia stem-cell development. Nat. Rev. Cancer. 2007;7:823–833. doi: 10.1038/nrc2253. PubMed DOI
Zhang L., Chen F., Zhang Z., Chen Y., Lin Y., Wang J. Design, synthesis and evaluation of the multidrug resistance-reversing activity of pyridine acid esters of podophyllotoxin in human leukemia cells. Bioorg. Med. Chem. Lett. 2016;26:4466–4471. doi: 10.1016/j.bmcl.2016.07.072. PubMed DOI
Lee K.-I., Su C.-C., Yang C.-Y., Hung D.-Z., Lin C.-T., Lu T.-H., Liu S.-H., Huang C.-F. Etoposide induces pancreatic β-cells cytotoxicity via the JNK/ERK/GSK-3 signaling-mediated mitochondria-dependent apoptosis pathway. Toxicol. Vitro. 2016;36:142–152. doi: 10.1016/j.tiv.2016.07.018. PubMed DOI
Pellegrini G.G., Morales C.C., Wallace T.C., Plotkin L.I., Bellido T. Avenanthramides Prevent Osteoblast and Osteocyte Apoptosis and Induce Osteoclast Apoptosis in Vitro in an Nrf2-Independent Manner. Nutrients. 2016;8:423. doi: 10.3390/nu8070423. PubMed DOI PMC
Papież M.A., Krzyściak W., Szade K., Bukowska-Straková K., Kozakowska M., Hajduk K., Bystrowska B., Dulak J., Jozkowicz A. Curcumin enhances the cytogenotoxic effect of etoposide in leukemia cells through induction of reactive oxygen species. Drug Des. Dev. Ther. 2016;10:557–570. doi: 10.2147/DDDT.S92687. PubMed DOI PMC
Zhang S., Lu C., Zhang X., Li J., Jiang H. Targeted delivery of etoposide to cancer cells by folate-modified nanostructured lipid drug delivery system. Drug Deliv. 2016;23:1838–1845. doi: 10.3109/10717544.2016.1141258. PubMed DOI
Lindsay G.S., Wallace H.M. Changes in polyamine catabolism in HL-60 human promyelogenous leukaemic cells in response to etoposide-induced apoptosis. Biochem. J. 1999;337 Pt 1:83–87. doi: 10.1042/bj3370083. PubMed DOI PMC
Kumar A., Ehrenshaft M., Tokar E.J., Mason R.P., Sinha B.K. Nitric oxide inhibits topoisomerase II activity and induces resistance to topoisomerase II-poisons in human tumor cells. Biochim. Biophys. Acta. 2016;1860:1519–1527. doi: 10.1016/j.bbagen.2016.04.009. PubMed DOI PMC
Zhang A., Lyu Y.L., Lin C.-P., Zhou N., Azarova A.M., Wood L.M., Liu L.F. A protease pathway for the repair of topoisomerase II–DNA covalent complexes. J. Biol. Chem. 2006;281:35997–36003. doi: 10.1074/jbc.M604149200. PubMed DOI
Ledesma F.C., El Khamisy S.F., Zuma M.C., Osborn K., Caldecott K.W. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature. 2009;461:674–678. doi: 10.1038/nature08444. PubMed DOI
Aparicio T., Baer R., Gottesman M., Gautier J. MRN, CtIP, and BRCA1 mediate repair of topoisomerase II–DNA adducts. J. Cell Biol. 2016;212:399–408. doi: 10.1083/jcb.201504005. PubMed DOI PMC
Quennet V., Beucher A., Barton O., Takeda S., Löbrich M. CtIP and MRN promote non-homologous end-joining of etoposide-induced DNA double-strand breaks in G1. Nucleic Acids Res. 2011;39:2144–2152. doi: 10.1093/nar/gkq1175. PubMed DOI PMC
Adachi N., Suzuki H., Iiizumi S., Koyama H. Hypersensitivity of nonhomologous DNA end-joining mutants to VP-16 and ICRF-193: Implications for the repair of topoisomerase II-mediated DNA damage. J. Biol. Chem. 2003;278:35897–35902. doi: 10.1074/jbc.M306500200. PubMed DOI
Excessive reactive oxygen species induce transcription-dependent replication stress
Autophagy role(s) in response to oncogenes and DNA replication stress