Acessibilidade / Reportar erro

DNA replication stress: oncogenes in the spotlight

Abstract

Precise replication of genetic material is essential to maintain genome stability. DNA replication is a tightly regulated process that ensues faithful copies of DNA molecules to daughter cells during each cell cycle. Perturbation of DNA replication may compromise the transmission of genetic information, leading to DNA damage, mutations, and chromosomal rearrangements. DNA replication stress, also referred to as DNA replicative stress, is defined as the slowing or stalling of replication fork progression during DNA synthesis as a result of different insults. Oncogene activation, one hallmark of cancer, is able to disturb numerous cellular processes, including DNA replication. In fact, extensive work has indicated that oncogene-induced replication stress is an important source of genomic instability in human carcinogenesis. In this review, we focus on main oncogenes that induce DNA replication stress, such as RAS, MYC, Cyclin E, MDM2, and BCL-2 among others, and the molecular mechanisms by which these oncogenes interfere with normal DNA replication and promote genomic instability.

Keywords:
Cancer; cell cycle; DNA replication; oncogene; replication stress

DNA replication

Eukaryotic chromosomes are precisely replicated once each cell cycle to ensure genome stability. The process of DNA replication is conserved among different organisms and is tightly controlled by the sequential assembly of various proteins onto DNA replication origins (ORIs), followed by the concerted synthesis of nascent DNA strands. In mammalian cells, ORIs are generally characterized as nucleosome-free, GC-rich genomic regions where DNA replication starts. Multiple protein complexes function in a coordinated fashion to recognize ORIs, unwind double-strand DNA, and perform DNA synthesis. Through the renowned semiconservative process, DNA replication is performed by different DNA polymerases, which require single-strand DNA (ssDNA) templates to build complementary DNA molecules: one continuous strand in the same direction as the replication fork progression (the leading strand) and one discontinuous strand in the opposite direction through the generation of short Okazaki fragments (the lagging strand) (Masai et al., 2010Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N and Oda M (2010) Eukaryotic chromosome DNA replication: where, when, and how? Annu Rev Biochem 79:89-130.; Leonard and Méchali, 2013Leonard AC and Méchali M (2013) DNA replication origins. Cold Spring Harb Perspect Biol 5:a010116.; O’Donnell et al., 2013O’Donnell M, Langston L and Stillman B (2013) Principles and concepts of DNA replication in bacteria, archea, and eukarya. Cold Spring Harb Perspect Biol 5:a010108.).

To ensure one round of DNA replication per cell cycle, cells precisely control the execution of two temporally separated steps before the onset of DNA synthesis: origin licensing and origin firing. During late mitosis and early G1 phase, when cells experience low cyclin-dependent kinase (CDK) environments, origin licensing is accomplished by the sequential assembly of protein complexes onto ORIs. Origin licensing occurs through the loading of origin recognition complex subunits 1-6 (ORC1-6), cell division cycle 6 (CDC6) protein, and chromatin licensing and DNA replication factor 1 (CDT1), followed by recruitment of DNA helicase minichromosome maintenance complex components 2-7 (MCM2-7) to form pre-replication complexes (pre-RC). At the pre-RC stage, the helicase complex is inactive and unable to unwind the double-strand DNA molecule. Once origin licensing is completed, cells activate several mechanisms to inhibit a new round of origin licensing within the same cell cycle, and therefore prevent DNA rereplication, such as inhibitory phosphorylation and ubiquitin-mediated degradation of pre-RC components among other mechanisms (Masai et al., 2010Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N and Oda M (2010) Eukaryotic chromosome DNA replication: where, when, and how? Annu Rev Biochem 79:89-130.; McIntosh and Blow, 2012McIntosh D and Blow JJ (2012) Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb Perspect Biol 4:a012955.; Siddiqui et al., 2013Siddiqui K, On KF and Diffley JF (2013) Regulating DNA replication in eukarya. Cold Spring Harb Perspect Biol 5:a012930.).

The second critical step before the onset of DNA replication occurs during the G1/S phase transition, when additional proteins are assembled onto chromatin to establish pre-initiation complexes (pre-IC). Contrary to origin licensing, origin activation requires high CDK activity and is triggered by the concerted action of CDC7 and CDK2 protein kinases, which associate with the regulatory subunits DBF4 and Cyclin E/A, respectively. These S phase kinases phosphorylate several replication factors during pre-IC assembly and activate the DNA helicase complex through facilitating the recruitment of CDC45 and GINS complex to form the CMG complex (CDC45-MCM-GINS). Activation of the CMG helicase then unwinds the double-strand DNA and further allows the recruitment of other replication factors, such as replication factor C (RFC), replication protein A (RPA), the sliding clamp proliferating cell nuclear antigen (PCNA), and multiple DNA polymerases, all essential for initiation of DNA synthesis and replication fork movement (replisome formation). It is important to point out that the vast majority of licensed origins along the genome are not activated during normal S phases and remain on hold as backup (dormant) ORIs to serve in specific physiological situations, such as DNA replication stress. Furthermore, the subset of activated origins in a given cell varies at each cell cycle and also differs among different cells, underscoring the importance of ORI activation dynamics and flexibility in DNA replication and other cellular functions (Masai et al., 2010Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N and Oda M (2010) Eukaryotic chromosome DNA replication: where, when, and how? Annu Rev Biochem 79:89-130.; Méchali, 2010Méchali M (2010) Eukaryotic DNA replication origins: Many choices for appropriate answers. Nat Rev Mol Cell Biol 11:728-738.; Tanaka and Araki, 2013Tanaka S and Araki H (2013) Helicase activation and establishment of replication forks at chromosomal origins of replication. Cold Spring Harb Perspect Biol 5:a010371.; Fragkos et al., 2015Fragkos M, Ganier O, Coulombe P and Méchali M (2015) DNA replication origin activation in space and time. Nat Rev Mol Cell Biol 16:360-374.).

Once ORIs are activated, DNA synthesis is triggered in S phase by replisomes (large replication machineries) at thousands of chromosomal sites with two replication forks progressing in opposite directions, a process known as origin firing. In close association with several replication factors (such as TopBP1, RecQL4, Treslin, and MCM10), the CMG complex moves along the DNA molecule, generating transient ssDNA and replication forks. DNA polymerases then catalyze the incorporation of deoxyribonucleoside triphosphates (dNTPs) to build two DNA strands that are complementary to the parental DNA molecule. DNA replication priming (synthesis initiation of a new DNA strand) is accomplished by the DNA polymerase alpha-primase complex, which synthesizes RNA/DNA hybrid primers, while replication elongation is primarily performed by DNA polymerase epsilon at the leading strand and DNA polymerase delta at the lagging strand through generation of 100-200 nucleotides long Okazaki fragments. In normal cell cycles, origin firing occurs at approximately 30-50,000 sites along the 3 billion base pairs of human chromosomes and DNA replication forks travel roughly at 1-2 Kb per minute, ensuring completion of chromosomal replication in about 8 hours during S phase. Importantly, DNA polymerases exonucleolytic proofreading activities and sophisticated DNA repair mechanisms work in coordination to generate high fidelity DNA molecules and preserve genome integrity (Johansson and Dixon, 2013Johansson E and Dixon N (2013) Replicative DNA polymerases. Cold Spring Harb Perspect Biol 5:a012799.; Lujan et al., 2016Lujan SA, Williams JS and Kunkel TA (2016) DNA polymerases divide the labor of genome replication. Trends Cell Biol 26:640-654.; Burgers and Kunkel, 2017Burgers PMJ and Kunkel TA (2017) Eukaryotic DNA replication fork. Annu Rev Biochem 86:417-438.).

Mechanisms of oncogene-induced replication stress

DNA replication stress, also known as DNA replicative stress, is characterized by the slowing or stalling of replication fork progression during DNA synthesis, which may lead to replication fork collapse and DNA damage. If not resolved by replication checkpoint mechanisms, persistent replication stress may cause mutations, copy number alterations (CNAs, amplifications and deletions), and chromosomal rearrangements (Zeman and Cimprich, 2014Zeman MK and Cimprich KA (2014) Causes and consequences of replication stress. Nat Cell Biol 16:2-9.; Gaillard et al., 2015Gaillard H, García-Muse T and Aguilera A (2015) Replication stress and cancer. Nat Rev Cancer 15:276-289.; Técher et al., 2017Técher H, Koundrioukoff S, Nicolas A and Debatisse M (2017) The impact of replication stress on replication dynamics and DNA damage in vertebrate cells. Nat Rev Genet 18:535-550.). In normal conditions, one of the main consequences of DNA replication stress is the activation of the DNA damage response (DDR) pathway, which is primarily triggered by the generation of ssDNA upon fork stalling. ssDNA creates a platform for recruitment and activation of several proteins, such as RPA, Ataxia Telangiectasia and Rad3-related (ATR), and Checkpoint Kinase 1 (CHK1), which subsequently recruit and activate numerous substrates to inhibit cell cycle progression, stabilize stalled replication forks, and promote DNA replication restart. Importantly, activation of the DDR pathway has been proposed to function as an inducible barrier during early stages of tumorigenesis, leading to cell cycle arrest, cell death or senescence. DDR deficiency compromises cellular checkpoints, causes DNA damage, and genomic instability, and is associated with cancer susceptibility (Bartkova et al., 2005Bartkova J, Horejsí Z, Koed K, Kramer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C et al. (2005) DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434:864-870., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-; Gorgoulis et al., 2005Gorgoulis VG, Vassiliou LV, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, Venere M, DiTullio RA Jr, Kastrinakis NG, Levy B et al. (2005) Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434:907-913.; Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.; Halazonetis et al., 2008Halazonetis TD, Gorgoulis VG and Bartek J (2008) An oncogene-induced DNA damage model for cancer development. Science 319:1352-1355.). The mechanisms of DDR activation upon DNA replication stress have been extensively reviewed in the literature and are beyond the scope of this article (Sirbu and Cortez, 2013Sirbu BM and Cortez D (2013) DNA damage response: Three levels of DNA repair regulation. Cold Spring Harb Perspect Biol 5:a012724.; Blackford and Jackson, 2017Blackford AN and Jackson SP (2017) ATM, ATR, and DNA-PK: The trinity at the heart of the DNA damage response. Mol Cell 66:801-817.; Saldivar et al., 2017Saldivar JC, Cortez D and Cimprich KA (2017) The essential kinase ATR: Ensuring faithful duplication of a challenging genome. Nat Rev Mol Cell Biol 18:622-636.; Toledo et al., 2017Toledo L, Neelsen KJ and Lukas J (2017) Replication catastrophe: when a checkpoint fails because of exhaustion. Mol Cell 66:735-749.). In this section, we briefly discuss the main mechanisms of oncogene-induced replication stress.

Oncogene activation, one established hallmark of cancer, is able to directly interfere with normal DNA replication and represents an important source of replication stress and genomic instability. Oncogene activation causes replication stress through different mechanisms, such as impairment of origin licensing and/or origin firing, nucleotide pool depletion, and interference between DNA replication and transcription machineries (Figure 1).

Figure 1
Molecular mechanisms of DNA replication stress. A) Unusual DNA secondary structures may be formed at certain genomic regions, such as centromeres, telomeres, and fragile sites, and represent natural obstacles to replication fork progression. Stem-loop (left and right) and G-quadruplex (middle) structures are represented. B) Impaired origin licensing may compromise the formation of active replication origins and DNA replication. Normal (left), impaired (middle), and absence of (right) pre-RC formation are represented. C) Disturbed origin firing may interfere with DNA replication and replication fork progression. Normal (right), asymmetric (middle), and repetitive (left) origin firing are represented. D) Uncontrolled S phase entry in the presence of nucleotide pool depletion may impair DNA replication and prevent replication fork progression. E) Collisions between replication and transcription machineries may impair DNA replication fork progression through generation of DNA topological stress and formation of persistent R-loops, RNA-DNA hybrid molecules. A-E) DNA molecule (blue strand), DNA origin of replication (Origin, red strand), ORC complex (green), CDC6 protein (orange), CDT1 protein (purple), MCM complex (yellow), RNA polymerase (blue), and messenger RNA (yellow strand) are represented. DNA polymerases and replisomes are omitted for simplicity. Grey arrows represent progression of replication or transcription machineries and red crosses represent stalled replication forks.

Unusual DNA structures may be formed at specific genomic regions during certain cellular processes that generate ssDNA, such as DNA replication, transcription, and DNA repair (Bochman et al., 2012Bochman ML, Paeschke K and Zakian VA (2012) DNA secondary structures: stability and function of G-quadruplex structures. Nat Rev Genet 13:770-780.; Kaushal and Freudenreich, 2019Kaushal S and Freudenreich CH (2019) The role of fork stalling and DNA structures in causing chromosome fragility. Genes Chromosomes Cancer 58:270-283.). Formation of DNA secondary structures normally occurs at repetitive nucleotide sequences and represents one important obstacle to replisome progression (Figure 1A). Several different alternative DNA structures, such as stem-loop and G-quadruplex (G4), may be formed at AT- and GC-rich regions, and can lead to increased DNA torsional stress, replication fork stalling, double-strand DNA breaks (DSBs), and chromosome fragility (Ozeri-Galai et al., 2011Ozeri-Galai E, Lebofsky R, Rahat A, Bester AC, Bensimon A and Kerem B (2011) Failure of origin activation in response to fork stalling leads to chromosomal instability at fragile sites. Mol Cell 43:122-131.; Chambers et al., 2015Chambers VS, Marsico G, Boutell JM, Di Antonio M, Smith GP and Balasubramanian S (2015) High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat Biotechnol 33:877-881.; Tubbs et al., 2018Tubbs A, Sridharan S, van Wietmarschen N, Maman Y, Callen E, Stanlie A, Wu W, Wu X, Day A, Wong N et al. (2018) Dual roles of poly(dA:dT) tracts in replication initiation and fork collapse. Cell 174:1127-1142.). In fact, oncogene activation may interfere with normal replication and pose further risk to genomic regions with these DNA secondary structures, which have been mapped to breakpoint hotspots and regions with CNAs in human cancers (Tsantoulis et al., 2008Tsantoulis PK, Kotsinas A, Sfikakis PP, Evangelou K, Sideridou M, Levy B, Mo L, Kittas C, Wu XR, Papavassiliou AG et al. (2008) Oncogene-induced replication stress preferentially targets common fragile sites in preneoplastic lesions. A genome-wide study. Oncogene 27:3256-3264.; Beroukhim et al., 2010Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, Donovan J, Barretina J, Boehm JS, Dobson J, Urashima M et al. (2010) The landscape of somatic copy-number alteration across human cancers. Nature 463:899-905.; Bignell et al., 2010Bignell GR, Greenman CD, Davies H, Butler AP, Edkins S, Andrews JM, Buck G, Chen L, Beare D, Latimer C et al. (2010) Signatures of mutation and selection in the cancer genome. Nature 463:893-898.). The consequences of unusual DNA structures to chromosome replication and fragility will be further discussed in the next section.

Origin licensing is the initial step of DNA replication and must be precisely coordinated through the cell cycle to allow appropriate origin firing in S phase (McIntosh and Blow, 2012McIntosh D and Blow JJ (2012) Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb Perspect Biol 4:a012955.). As discussed before, the vast majority of licensed origins constitute backup (dormant) ORIs that are not activated during normal S phase. Accordingly, it has been shown that depletion of pre-RC proteins does not interfere with DNA replication in unperturbed cells (Ge et al., 2007Ge XQ, Jackson DA and Blow JJ (2007) Dormant origins licensed by excess Mcm2-7 are required for human cells to survive replicative stress. Genes Dev 21:3331-3341.; Ibarra et al., 2008Ibarra A, Schowb E and Méndez J (2008) Excess MCM proteins protect human cells from replicative stress by licensing backup origins of replication. Proc Natl Acad Sci U S A 105:8956-8961.). However, under conditions of challenged DNA replication, deficient assembly of pre-RC proteins reduces the number of functional ORIs, impairing DNA replication and causing replication stress (Figure 1B). Indeed, substantial interference with ORC2, CDT1 or MCM2 loading onto chromatin arrests cells in G1 and prevents S phase progression, most likely because of insufficient origin licensing (Shreeram et al., 2002Shreeram S, Sparks A, Lane DP and Blow JJ (2002) Cell type-specific responses of human cells to inhibition of replication licensing. Oncogene 21:6624-6632.; Machida et al., 2005Machida YJ, Teer JK and Dutta A (2005) Acute reduction of an origin recognition complex (ORC) subunit in human cells reveals a requirement of ORC for Cdk2 activation. J Biol Chem 280:27624-27630.). Oncogene activation has also been shown to directly inhibit the loading of MCM complex proteins onto chromatin, resulting in impaired origin firing and fork progression (Ekholm-Reed et al., 2004Ekholm-Reed S, Méndez J, Tedesco D, Zetterberg A, Stillman B and Reed SI (2004) Deregulation of cyclin E in human cells interferes with prereplication complex assembly. J Cell Biol 165:789-800.; Bartkova et al., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-).

Following origin licensing, coordinated origin firing is also essential for accurate DNA replication (Fragkos et al., 2015Fragkos M, Ganier O, Coulombe P and Méchali M (2015) DNA replication origin activation in space and time. Nat Rev Mol Cell Biol 16:360-374.). Reduced or asymmetric origin firing may force replication forks to travel for longer distances along the genome, increasing the chances of replication fork collapse (Figure 1C). Impaired ORI activation may also decrease replication fork velocity, allowing cells to enter into mitosis with incompletely replicated genomes. In fact, oncogene activation has been shown to inhibit origin firing and lead to unscheduled DNA replication (Frum et al., 2014Frum RA, Singh S, Vaughan C, Mukhopadhyay ND, Grossman SR, Windle B, Deb S and Deb SP (2014) The human oncoprotein MDM2 induces replication stress eliciting early intra-S-phase checkpoint response and inhibition of DNA replication origin firing. Nucleic Acids Res 42:926-940.). On the other hand, oncogene activation may also induce replication stress through increased origin firing (Vaziri et al., 2003Vaziri C, Saxena S, Jeon Y, Lee C, Murata K, Machida Y, Wagle N, Hwang DS and Dutta A (2003) A p53-dependent checkpoint pathway prevents rereplication. Mol Cell 11:997-1008.). Multiple ORI activation at specific genomic sites can lead to a second round of DNA replication within one cell cycle, a process known as DNA rereplication (Figure 1C). Indeed, overexpression of pre-RC components, such as CDT1 and CDC6, increases origin firing, induces DNA rereplication, and has been observed in different human cancers (Vaziri et al., 2003Vaziri C, Saxena S, Jeon Y, Lee C, Murata K, Machida Y, Wagle N, Hwang DS and Dutta A (2003) A p53-dependent checkpoint pathway prevents rereplication. Mol Cell 11:997-1008.; Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.; Liontos et al., 2007Liontos M, Koutsami M, Sideridou M, Evangelou K, Kletsas D, Levy B, Kotsinas A, Nahum O, Zoumpourlis V, Kouloukoussa M et al. (2007) Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res 67:10899-10909.).

Nucleotides are essential components of nucleic acids and are necessary for DNA replication (Lane and Fan, 2015Lane AN and Fan TW (2015) Regulation of mammalian nucleotide metabolism and biosynthesis. Nucleic Acids Res 43:2466-2485.). The nucleotide biosynthesis pathway must be precisely coordinated within cells to maintain normal levels of deoxyribonucleotides and ensure normal DNA replication. Oncogene activation may induce uncontrolled S phase entry with insufficient nucleotide pools (Figure 1D). In fact, it has been shown that oncogene overexpression is able to induce increased cell proliferation with exhausted dNTP levels, leading to replication fork stalling and DSBs (Bester et al., 2011Bester AC, Roniger M, Oren YS, Im MM, Sarni D, Chaoat M, Bensimon A, Zamir G, Shewach DS and Kerem B (2011) Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145:435-446.). Also, oncogene activation may directly interfere with nucleotide biosynthesis, causing dNTP pool depletion and premature termination of replication forks (Aird et al., 2013Aird KM, Zhang G, Li H, Tu Z, Bitler BG, Garipov A, Wu H, Wei Z, Wagner SN, Herlyn M et al. (2013) Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep 3:1252-1265.; Xie et al., 2013Xie M, Yen Y, Owonikoko TK, Ramalingam SS, Khuri FR, Curran WJ, Doetsch PW and Deng X (2013) Bcl2 induces DNA replication stress by inhibiting ribonucleotide reductase. Cancer Res 74:212-223.).

Finally, DNA replication stress may be also caused by collisions between replication and transcription machineries. These conflicts usually occur at genomic sites that encode large genes (> 800 Kb), which require more than one round of cell cycle to complete transcription and therefore are transcriptionally active during S phase (Helmrich et al., 2011Helmrich A, Ballarino M and Tora L (2011) Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol Cell 44:966-977.). Transcription-replication collisions may lead to DNA topological constraints and persistent accumulation of R-loops, RNA-DNA hybrid molecules generated during transcription (Helmrich et al., 2013Helmrich A, Ballarino M, Nudler E and Tora L (2013) Transcription-replication encounters, consequences and genomic instability. Nat Struct Mol Biol 20:412-418.). If not resolved, these structures may cause replication fork stalling, DNA damage, and chromosome breakage (Figure 1E). Another potential consequence of unresolved transcription-replication collisions is the formation of unusual DNA replication intermediates, such as reversed replication forks (Neelsen and Lopes, 2015Neelsen KJ and Lopes M (2015) Replication fork reversal in eukaryotes: from dead end to dynamic response. Nat Rev Mol Cell Biol 16:207-220.). Indeed, it has been shown that oncogene activation induces conflicts between replication and transcription machineries due to increased transcriptional activity and R-loop formation, leading to replication stress and DNA damage (Jones et al., 2013Jones RM, Mortusewicz O, Afzal I, Lorvellec M, García P, Helleday T and Petermann E (2013) Increased replication initiation and conflicts with transcription underlie Cyclin E-induced replication stress. Oncogene 32:3744-3753.; Kotsantis et al., 2016Kotsantis P, Silva LM, Irmscher S, Jones RM, Folkes L, Gromak N and Petermann E (2016) Increased global transcription activity as a mechanism of replication stress in cancer. Nat Commun 7:13087.). The molecular mechanisms of oncogene-induced replication stress have been discussed in detail by others (Hills and Diffley, 2014Hills SA and Diffley JF (2014) DNA replication and oncogene-induced replicative stress. Curr Biol 24:R435-R444.; Macheret and Halazonetis, 2015Macheret M and Halazonetis TD (2015) DNA replication stress as a hallmark of cancer. Annu Rev Pathol 10:425-448.; Kotsantis et al., 2018Kotsantis P, Petermann E and Boulton SJ (2018) Mechanisms of oncogene-induced replication stress: Jigsaw falling into place. Cancer Discov 8:537-555.).

Genomic regions susceptible to replication stress

Certain genomic regions present intrinsic difficulties to accomplish DNA synthesis upon perturbed DNA replication. Among these regions, common fragile sites (CFS) have been defined as chromosomal loci that are prone to breaks and/or gaps in situations of replication stress. These sites are usually characterized by AT-rich sequences and ORI paucity, are located at late-replicating domains, and contain large isolated genes (Debatisse et al., 2012Debatisse M, Le Tallec B, Letessier A, Dutrillaux B and Brison O (2012) Common fragile sites: Mechanisms of instability revisited. Trends Genet 28:22-32.; Ozeri-Galai et al., 2012Ozeri-Galai E, Bester AC and Kerem B (2012) The complex basis underlying common fragile site instability in cancer. Trends Genet 28:295-302.; Glover et al., 2017Glover TW, Wilson TE and Arlt MF (2017) Fragile sites in cancer: More than meets the eye. Nat Rev Cancer 17:489-501.). Repetitive AT sequences may lead to formation of DNA secondary structures, which impose natural obstacles to replication fork progression (Ozeri-Galai et al., 2011Ozeri-Galai E, Lebofsky R, Rahat A, Bester AC, Bensimon A and Kerem B (2011) Failure of origin activation in response to fork stalling leads to chromosomal instability at fragile sites. Mol Cell 43:122-131.). Lack of ORI activation forces distant converging replication forks to travel for long distances to finish DNA synthesis, increasing the risk of incomplete DNA replication (Letessier et al., 2011Letessier A, Millot GA, Koundrioukoff S, Lachagès AM, Vogt N, Hansen RS, Malfoy B, Brison O and Debatisse M (2011) Cell-type-specific replication initiation programs set fragility of the FRA3B fragile site. Nature 470:120-123.). Genomic regions that replicate late in S phase also present an increased likelihood of incomplete DNA replication because there might not be enough time to complete DNA synthesis within S phase (Le Beau et al., 1998Le Beau MM, Rassool FV, Neilly ME, Espinosa R 3rd, Glover TW, Smith DI and McKeithan TW (1998) Replication of a common fragile site, FRA3B, occurs late in S phase and is delayed further upon induction: implications for the mechanism of fragile site induction. Hum Mol Genet 7:755-761.). Finally, as discussed before, chromosomal loci with large, actively transcribed genes are more susceptible to collisions between replication and transcription machineries, also contributing to CFS instability (Helmrich et al., 2011Helmrich A, Ballarino M and Tora L (2011) Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol Cell 44:966-977.).

CFS strongly correlate with recurrent deletions in a broad spectrum of human tumors (Tsantoulis et al., 2008Tsantoulis PK, Kotsinas A, Sfikakis PP, Evangelou K, Sideridou M, Levy B, Mo L, Kittas C, Wu XR, Papavassiliou AG et al. (2008) Oncogene-induced replication stress preferentially targets common fragile sites in preneoplastic lesions. A genome-wide study. Oncogene 27:3256-3264.; Beroukhim et al., 2010Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, Donovan J, Barretina J, Boehm JS, Dobson J, Urashima M et al. (2010) The landscape of somatic copy-number alteration across human cancers. Nature 463:899-905.; Bignell et al., 2010Bignell GR, Greenman CD, Davies H, Butler AP, Edkins S, Andrews JM, Buck G, Chen L, Beare D, Latimer C et al. (2010) Signatures of mutation and selection in the cancer genome. Nature 463:893-898.). FRA3B and FRA16D are the two most frequently affected CFS in human cancers, including breast, lung, colon, esophageal, and renal carcinomas (Durkin and Glover, 2007Durkin SG and Glover TW (2007) Chromosome fragile sites. Annu Rev Genet 41:169-192.). FRA3B is located at 3p14.2 and overlaps with the 1.5 Mb-long Fragile Histidine Triad (FHIT) tumor suppressor gene, which is involved in nucleotide metabolism (Saldivar and Park, 2019Saldivar JC and Park D (2019) Mechanisms shaping the mutational landscape of the FRA3B/FHIT-deficient cancer genome. Genes Chromosomes Cancer 58:317-323.). FRA3B instability is caused by a paucity of replication initiation events at the central region of this fragile site, as well as transcription-replication collisions due to extended transcription of the large FHIT gene (Helmrich et al., 2011Helmrich A, Ballarino M and Tora L (2011) Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol Cell 44:966-977.; Letessier et al., 2011Letessier A, Millot GA, Koundrioukoff S, Lachagès AM, Vogt N, Hansen RS, Malfoy B, Brison O and Debatisse M (2011) Cell-type-specific replication initiation programs set fragility of the FRA3B fragile site. Nature 470:120-123.). FRA16D is located at 16q23 and overlaps with the 1.1 Mb-long WW Domain Containing Oxidoreductase (WWOX) tumor suppressor gene, which is involved in apoptotic and DDR pathways (Hussain et al., 2019Hussain T, Liu B, Shrock MS, Williams T and Aldaz CM (2019) WWOX, the FRA16D gene: A target of and a contributor to genomic instability. Genes Chromosomes Cancer 58:324-338.). Similar to FRA3B, FRA16D fragility is also associated with scarcity of initiation events and transcription-replication collisions at the large WWOX gene (Helmrich et al., 2011Helmrich A, Ballarino M and Tora L (2011) Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol Cell 44:966-977.; Letessier et al., 2011Letessier A, Millot GA, Koundrioukoff S, Lachagès AM, Vogt N, Hansen RS, Malfoy B, Brison O and Debatisse M (2011) Cell-type-specific replication initiation programs set fragility of the FRA3B fragile site. Nature 470:120-123.). In addition to FRA3B and FRA16D, other CFS, such as FRA6E, FRA9E, and FRA7G, present intrinsic vulnerabilities, are susceptible to major genomic losses, and have been shown to contribute to human carcinogenesis (Durkin and Glover, 2007Durkin SG and Glover TW (2007) Chromosome fragile sites. Annu Rev Genet 41:169-192.; Glover et al., 2017Glover TW, Wilson TE and Arlt MF (2017) Fragile sites in cancer: More than meets the eye. Nat Rev Cancer 17:489-501.).

Although late-replicating genomic regions are susceptible to chromosomal fragility, early-replicating fragile sites (ERFS) have also been shown to be vulnerable to replication stress and DNA damage (Mortusewicz et al., 2013Mortusewicz O, Herr P and Helleday T (2013) Early replication fragile sites: Where replication-transcription collisions cause genetic instability. EMBO J 32:493-495.). Unlike CFS, ERFS are characterized by GC-rich sequences, repetitive elements, increased ORI density, and highly transcribed gene clusters. Upon S phase entry, these genomic regions show high ORI activity close to transcriptionally active genes, leading to replication fork stalling, DSBs, and chromosome rearrangements (Barlow et al., 2013Barlow JH, Faryabi RB, Callén E, Wong N, Malhowski A, Chen HT, Gutierrez-Cruz G, Sun HW, McKinnon P, Wright G et al. (2013) Identification of early replicating fragile sites that contribute to genome instability. Cell 152:620-632.). It is therefore likely that ERFS instability is induced by increased conflicts between replication and transcription machineries. Importantly, many ERFS overlap with recurrent CNAs at genomic regions implicated in the development of human diffuse large B cell lymphomas (Barlow et al., 2013Barlow JH, Faryabi RB, Callén E, Wong N, Malhowski A, Chen HT, Gutierrez-Cruz G, Sun HW, McKinnon P, Wright G et al. (2013) Identification of early replicating fragile sites that contribute to genome instability. Cell 152:620-632.).

Besides CFS and ERFS, other genomic regions are also inherently difficult to replicate and susceptible to replication stress. Two clear examples are telomeres and centromeres, which are both heterochromatic regions enriched in repetitive sequences. These chromosomal regions are prone to formation of complex DNA secondary structures, such as stem-loops, G4 structures, and DNA catenanes, which can interfere with replication fork progression and contribute to chromosome fragility (Martínez and Blasco, 2015Martínez P and Blasco MA (2015) Replicating through telomeres: A means to an end. Trends Biochem Sci 40:504-515.; Bloom and Costanzo, 2017Bloom K and Costanzo V (2017) Centromere structure and function. In: Black BE (ed) Centromeres and kinetochores. Springer, Cham, pp 515-539.; Higa et al., 2017Higa M, Fujita M and Yoshida K (2017) DNA replication origins and fork progression at mammalian telomeres. Genes 8:E112.; Black and Giunta, 2018Black EM and Giunta S (2018) Repetitive fragile sites: Centromere satellite DNA as a source of genome instability in human diseases. Genes 9:E615.). Sophisticated protein complexes regulate telomere and centromere stability and function. Disruption of several telomere- and centromere-binding proteins has been shown to impair resolution of DNA secondary structures, induce replication fork stalling, and cause fragility at these loci (Martínez et al., 2009Martínez P, Thanasoula M, Muñoz P, Liao C, Tejera A, McNees C, Flores JM, Fernández-Capetillo O, Tarsounas M and Blasco MA (2009) Increased telomere fragility and fusions resulting from TRF1 deficiency lead to degenerative pathologies and increased cancer in mice. Genes Dev 23:2060-2075.; Sfeir et al., 2009Sfeir A, Kosiyatrakul ST, Hockemeyer D, MacRae SL, Karlseder J, Schildkraut CL and de Lange T (2009) Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138:90-103.; Aze et al., 2016Aze A, Sannino V, Soffientini P, Bachi A and Costanzo V (2016) Centromeric DNA replication reconstitution reveals DNA loops and ATR checkpoint suppression. Nat Cell Biol 18:684-691.; Giunta and Funabiki, 2017Giunta S and Funabiki H (2017) Integrity of the human centromere DNA repeats is protected by CENP-A, CENP-C, and CENP-T. Proc Natl Acad Sci U S A 114:1928-1933.). In addition, oncogene activation has been demonstrated to induce chromosome breaks at centromeres and generate aberrant structures at telomeres in response to replication stress (Suram et al., 2012Suram A, Kaplunov J, Patel PL, Ruan H, Cerutti A, Boccardi V, Fumagalli M, Di Micco R, Mirani N, Gurung RL et al. (2012) Oncogene-induced telomere dysfunction enforces cellular senescence in human cancer precursor lesions. EMBO J 31:2839-2851.; Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.).

Oncogenes in the spotlight

DNA replication must be precisely regulated during cell cycle in order to ensure genome stability. An extensive body of work has clearly demonstrated that oncogene activation induces replication stress at susceptible genomic sites through different molecular mechanisms (Figure 1). In the following sections, we discuss in detail the effects of the main human oncogenes that have been shown to cause DNA replication stress.

RAS

Oncogenic RAS has been closely related to DNA replication stress. The RAS family is composed of three proto-oncogenes (K-, H-, and N-RAS) that function as small GTPase signal transducers. RAS proteins are essential components of a network that communicate cell surface receptors with intracellular proteins to regulate cellular growth, survival, and metabolism among other functions. Under physiological conditions, these G proteins are activated upon GTP binding and then activate downstream effectors that regulate several mitogenic pathways, including the RAF/MEK/ERK and the PI3K/AKT pathways. Somatic mutations in RAS cause its constitutive activation and the subsequent stimulation of effectors that promote cell proliferation, apoptosis suppression, and metabolic reprogramming. RAS alterations are frequently observed in human cancers, specifically K-RAS mutations, which are found in approximately 40% of colorectal cancers and 20% of lung adenocarcinomas (Karnoub and Weinberg, 2008Karnoub AE and Weinberg RA (2008) Ras oncogenes: Split personalities. Nat Rev Mol Cell Biol 9:517-531.; Pylayeva-Gupta et al., 2011Pylayeva-Gupta Y, Grabocka E and Bar-Sagi D (2011) RAS oncogenes: Weaving a tumorigenic web. Nat Rev Cancer 11:761-774.).

Sustained mitogenic stimulation by oncogenic RAS (H-RASV12) directly impinges on DNA replication and causes replication stress through several mechanisms (Table 1). In a groundbreaking work, Di Micco and collaborators have shown that oncogenic RAS induces replication stress by increasing origin firing and generating asymmetric replication forks (Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.). It is possible that the increased origin firing reflects on DNA rereplication induced by the licensing factor CDC6, as it has been shown that RAS overexpression upregulates the levels of CDC6. It has also been demonstrated that oncogenic RAS interferes with cellular dNTP levels by downregulating the ribonucleotide reductase subunit M2 (RRM2), causing dNTP pool depletion and premature termination of replication forks (Aird et al., 2013Aird KM, Zhang G, Li H, Tu Z, Bitler BG, Garipov A, Wu H, Wei Z, Wagner SN, Herlyn M et al. (2013) Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep 3:1252-1265.). Together with others, these observations have contributed to the notion that oncogene-induced replication stress leads to a robust DDR activation and an irreversible cell cycle arrest, a phenotype known as oncogene-induced senescence (OIS) (Bartkova et al., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-; Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642., 2007Di Micco R, Fumagalli M and d’Adda di Fagagna F (2007) Breaking news: High-speed race ends in arrest - how oncogenes induce senescence. Trends Cell Biol 17:529-536.; Mallette et al., 2007Mallette FA, Gaumont-Leclerc MF and Ferbeyre G (2007) The DNA damage signaling pathway is a critical mediator of oncogene-induced senescence. Genes Dev 21:43-48.). In fact, oncogene-induced DDR activation, followed by cell death or senescence, has been proposed to function as an inducible barrier against human tumorigenesis (Bartkova et al., 2005Bartkova J, Horejsí Z, Koed K, Kramer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C et al. (2005) DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434:864-870., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-; Gorgoulis et al., 2005Gorgoulis VG, Vassiliou LV, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, Venere M, DiTullio RA Jr, Kastrinakis NG, Levy B et al. (2005) Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434:907-913.; Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.; Halazonetis et al., 2008Halazonetis TD, Gorgoulis VG and Bartek J (2008) An oncogene-induced DNA damage model for cancer development. Science 319:1352-1355.).

Table 1
Mechanisms of DNA replication stress induced by different oncogenes.

Oncogenic RAS may also induce replication stress as a consequence of oxidative stress. Initial expression of oncogenic RAS causes hyperproliferation and increases the velocity of replication forks. However, overexpression of RAS for longer periods of time causes cellular metabolic changes and reduces fork progression (Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.; Maya-Mendoza et al., 2015Maya-Mendoza A, Ostrakova J, Kosar M, Hall A, Duskova P, Mistrik M, Merchut-Maya JM, Hodny Z, Bartkova J, Christensen C et al. (2015) Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol Oncol 9:601-616.). It has been demonstrated that RAS-induced senescence is triggered by increased production of reactive oxygen species (ROS) (Irani et al., 1997Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, Fearon ER, Sundaresan M, Finkel T and Goldschmidt-Clermont PJ (1997) Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275:1649-1652.; Lee et al., 1999Lee AC, Fenster BE, Ito H, Takeda K, Bae NS, Hirai T, Yu ZX, Ferrans VJ, Howard BH and Finkel T (1999) Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J Biol Chem 274:7936-7940.), which lead to nucleotide oxidation as well as H2O2 generation (Rai et al., 2011Rai P, Young JJ, Burton DG, Giribaldi MG, Onder TT and Weinberg RA (2011) Enhanced elimination of oxidized guanine nucleotides inhibits oncogenic RAS-induced DNA damage and premature senescence. Oncogene 30:1489-1496.; Weyemi et al., 2012Weyemi U, Lagente-Chevallier O, Boufraqech M, Prenois F, Courtin F, Caillou B, Talbot M, Dardalhon M, Al Ghuzlan A, Bidart JM et al. (2012) ROS-generating NADPH oxidase NOX4 is a critical mediator in oncogenic H-Ras-induced DNA damage and subsequent senescence. Oncogene 31:1117-1129.). Alleviation of these oxidative insults by different approaches prevents DNA damage and cellular senescence. Therefore, it is possible that oxidative stress contributes to RAS-induced replication stress through accumulation of oxidized DNA precursors and generation of DSBs (Leikam et al., 2008Leikam C, Hufnagel A, Schartl M and Meierjohann S (2008) Oncogene activation in melanocytes links reactive oxygen to multinucleated phenotype and senescence. Oncogene 27:7070-7082.; Maya-Mendoza et al., 2015Maya-Mendoza A, Ostrakova J, Kosar M, Hall A, Duskova P, Mistrik M, Merchut-Maya JM, Hodny Z, Bartkova J, Christensen C et al. (2015) Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol Oncol 9:601-616.).

Another mechanism of replication stress induced by RAS is increased global transcription. RAS proteins promote cellular proliferation through upregulation of general transcription factors that are able to stimulate RNA synthesis (Pylayeva-Gupta et al., 2011Pylayeva-Gupta Y, Grabocka E and Bar-Sagi D (2011) RAS oncogenes: Weaving a tumorigenic web. Nat Rev Cancer 11:761-774.). Indeed, it has been shown that oncogenic RAS leads to elevated expression of the TBP transcription factor (TATA-box binding protein) and increased transcriptional activity. Elevated RNA synthesis causes replication fork slowing and DNA damage through collisions between replication and transcription machineries and subsequent formation of R-loops (Kotsantis et al., 2016Kotsantis P, Silva LM, Irmscher S, Jones RM, Folkes L, Gromak N and Petermann E (2016) Increased global transcription activity as a mechanism of replication stress in cancer. Nat Commun 7:13087.). Interestingly, TBP overexpression alone is able to increase transcription and cause replication stress and DNA damage, recapitulating the effects of oncogenic RAS.

Other mechanisms may also contribute to RAS-induced replication stress. One possibility is the interference with DNA repair. It has been shown that oncogenic RAS causes dissociation of BRCA1 protein from chromatin, compromising DNA repair and leading to DNA damage (Tu et al., 2011Tu Z, Aird KM, Bitler BG, Nicodemus JP, Beeharry N, Xia B, Yen TJ and Zhang R (2011) Oncogenic RAS regulates BRIP1 expression to induce dissociation of BRCA1 from chromatin, inhibit DNA repair, and promote senescence. Dev Cell 21:1077-1091.). Inactivation of BRCA1 protein renders cells susceptible to accumulation of secondary mutations and potentially cancer development. In light of the numerous insults caused by oncogenic RAS in DNA replication, it is reasonable to speculate that RAS-induced replication stress may result in genomic instability. In fact, RAS activation has been shown to induce chromosome abnormalities, such as acentric fragments, deletions, and double minute chromosomes (Denko et al., 1994Denko NC, Giaccia AJ, Stringer JR and Stambrook PJ (1994) The human Ha-ras oncogene induces genomic instability in murine fibroblasts within one cell cycle. Proc Natl Acad Sci U S A 91:5124-5128.; Guerra et al., 2003Guerra C, Mijimolle N, Dhawahir A, Dubus P, Barradas M, Serrano M, Campuzano V and Barbacid M (2003) Tumor incidence by and endogenous K-ras oncogene is highly dependent on cellular context. Cancer Cell 4:111-120.), replication fork stalling at telomeres, leading to telomere attrition and aberrant telomeric structures (Suram et al., 2012Suram A, Kaplunov J, Patel PL, Ruan H, Cerutti A, Boccardi V, Fumagalli M, Di Micco R, Mirani N, Gurung RL et al. (2012) Oncogene-induced telomere dysfunction enforces cellular senescence in human cancer precursor lesions. EMBO J 31:2839-2851.), and genomic alterations at CFS relevant to human carcinogenesis (Tsantoulis et al., 2008Tsantoulis PK, Kotsinas A, Sfikakis PP, Evangelou K, Sideridou M, Levy B, Mo L, Kittas C, Wu XR, Papavassiliou AG et al. (2008) Oncogene-induced replication stress preferentially targets common fragile sites in preneoplastic lesions. A genome-wide study. Oncogene 27:3256-3264.; Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.).

MYC

The MYC family of transcription factors in composed of the three members: C-, L-, and N-MYC. MYC proteins are effectors of several signaling transduction pathways and control a variety of cellular functions, including cell growth, proliferation, differentiation, and apoptosis. As a transcription factor, MYC primarily mediates its functions through dimerization with MAX and binding DNA regulatory elements to regulate an array of gene transcription programs. Additionally, MYC proteins also play nontranscriptional roles in cellular physiology. Activation of oncogenic MYC usually occurs through gene amplification, chromosomal rearrangement or loss of upstream MYC regulators, leading to sustained levels of MYC and interference with essential cellular processes. In fact, deregulation of c-MYC expression is observed in more than half of human cancers and oncogenic MYC has been associated with aggressive breast, prostate, and colon cancers, as well as Burkitt lymphoma (Dang, 2012Dang CV (2012) MYC on the path to cancer. Cell 149:22-35.; Dominguez-Sola and Gautier, 2014Dominguez-Sola D and Gautier J (2014) MYC and the control of DNA replication. Cold Spring Harb Perspect Med 4:a014423.; Rohban and Campaner, 2015Rohban S and Campaner S (2015) Myc induced replicative stress response: How to cope with it and exploit it. Biochim Biophys Acta 1849:517-254.).

MYC-induced replication stress is triggered by different molecular mechanisms and generates DNA damage and genomic instability during carcinogenesis (Table 1). Initial evidence indicated that MYC-induced genomic instability was associated with oxidative stress. c-MYC overexpression causes alterations in cellular metabolism, including increased production of ROS, which correlates with DNA damage (Vafa et al., 2002Vafa O, Wade M, Kern S, Beeche M, Pandita TK, Hampton GM and Wahl GM (2002) c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability. Mol Cell 9:1031-1044.). However, in contrast to RAS, oncogenic MYC causes replication stress before induction of cellular metabolic changes (Maya-Mendoza et al., 2015Maya-Mendoza A, Ostrakova J, Kosar M, Hall A, Duskova P, Mistrik M, Merchut-Maya JM, Hodny Z, Bartkova J, Christensen C et al. (2015) Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol Oncol 9:601-616.). In fact, several studies have subsequently demonstrated that MYC activation leads to DNA damage and genomic instability through direct impairment of DNA replication dynamics (Karlsson et al., 2003Karlsson A, Deb-Basu D, Cherry A, Turner S, Ford J and Felsher DW (2003) Defective double-strand DNA break repair and chromosomal translocations by MYC overexpression. Proc Natl Acad Sci U S A 100:9974-9979.; Ray et al., 2006Ray S, Atkuri KR, Deb-Basu D, Adler AS, Chang HY, Herzenberg LA and Felsher DW (2006) MYC can induce DNA breaks in vivo and in vitro independent of reactive oxygen species. Cancer Res 66:6598-6605.; Dominguez-Sola et al., 2007Dominguez-Sola D, Ying CY, Grandori C, Ruggiero L, Chen B, Li M, Galloway DA, Gu W, Gautier J and Dalla-Favera R (2007) Non-transcriptional control of DNA replication by c-Myc. Nature 448:445-451.; Sankar et al., 2009Sankar N, Kadeppagari RK and Thimmapaya B (2009) c-Myc-induced aberrant DNA synthesis and activation of DNA damage response in p300 knockdown cells. J Biol Chem 284:15193-15205.; Srinivasan et al., 2013Srinivasan SV, Dominguez-Sola D, Wang LC, Hyrien O and Gautier J (2013) Cdc45 is a critical effector of Myc-dependent DNA replication stress. Cell Rep 3:1629-1639.).

The main mechanism of MYC-induced replication stress is through interference with origin firing. It has been demonstrated that MYC localizes to ORIs and physically interacts with pre-RC components during origin licensing, such ORCs, CDC6, CDT1, and MCMs (Dominguez-Sola et al., 2007Dominguez-Sola D, Ying CY, Grandori C, Ruggiero L, Chen B, Li M, Galloway DA, Gu W, Gautier J and Dalla-Favera R (2007) Non-transcriptional control of DNA replication by c-Myc. Nature 448:445-451.). MYC also participates in ORI activation by increasing the recruitment of CDC45 to chromatin, a replication factor that is essential for initiation of DNA replication (Dominguez-Sola et al., 2007Dominguez-Sola D, Ying CY, Grandori C, Ruggiero L, Chen B, Li M, Galloway DA, Gu W, Gautier J and Dalla-Favera R (2007) Non-transcriptional control of DNA replication by c-Myc. Nature 448:445-451.; Srinivasan et al., 2013Srinivasan SV, Dominguez-Sola D, Wang LC, Hyrien O and Gautier J (2013) Cdc45 is a critical effector of Myc-dependent DNA replication stress. Cell Rep 3:1629-1639.). In accordance, MYC depletion decreases the number of active ORIs, while MYC overexpression leads to increased and premature origin firing. Once deregulated, oncogenic MYC leads to ORI hyperactivation, replication fork asymmetry and stalling, and eventually DNA damage (Dominguez-Sola et al., 2007Dominguez-Sola D, Ying CY, Grandori C, Ruggiero L, Chen B, Li M, Galloway DA, Gu W, Gautier J and Dalla-Favera R (2007) Non-transcriptional control of DNA replication by c-Myc. Nature 448:445-451.; Srinivasan et al., 2013Srinivasan SV, Dominguez-Sola D, Wang LC, Hyrien O and Gautier J (2013) Cdc45 is a critical effector of Myc-dependent DNA replication stress. Cell Rep 3:1629-1639.; Maya-Mendoza et al., 2015Maya-Mendoza A, Ostrakova J, Kosar M, Hall A, Duskova P, Mistrik M, Merchut-Maya JM, Hodny Z, Bartkova J, Christensen C et al. (2015) Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol Oncol 9:601-616.). Importantly, these effects of MYC on origin firing have been shown to be independent of its transcriptional activity. Similar to the well-characterized effect of oncogenic Cyclin E1 in origin firing (discussed below), MYC overexpression also induces changes in genomic location of ORI activation from intergenic to intragenic regions with high transcriptional activity (Macheret and Halazonetis, 2018Macheret M and Halazonetis TD (2018) Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555:112-116.). Considering that MYC is a transcription factor and that its overexpression upregulates transcription and increases origin firing, it is reasonable to speculate that oncogenic MYC also causes replication stress by generating collisions between replication and transcription machineries. However, this potential mechanism of MYC-mediated replication stress remains to be demonstrated.

An indirect mechanism for MYC-induced replication stress is through activation of Cyclin E/CDK2 complex. It has been widely demonstrated that oncogenic MYC promotes cell cycle progression and increases Cyclin E/CDK2 activity, which may be achieved by induction of CCND2 gene expression, inactivation of CDK inhibitor p27Kip1 or stimulation of E2F transcription factor-dependent genes, among other mechanisms (Bretones et al., 2015Bretones G, Delgado MD and León J (2015) Myc and cell cycle control. Biochim Biophys Acta 1849:506-516.). The specific consequences of increased Cyclin E/CDK2 activity to replication stress are discussed in the following section.

In contrast to the above, MYC proteins can intriguingly counteract replication stress through several mechanisms. As mentioned earlier, MYC transcription factors induce expression of numerous genes involved in cellular proliferation and DNA replication, including the nucleotide biosynthesis pathway (Liu et al., 2008Liu YC, Li F, Handler J, Huang CR, Xiang Y, Neretti N, Sedivy JM, Zeller KI and Dang CV (2008) Global regulation of nucleotide biosynthetic genes by c-Myc. PLoS One 3:e2722.; Mannava et al., 2008Mannava S, Grachtchouk V, Wheeler LJ, Im M, Zhuang D, Slavina EG, Mathews CK, Shewach DS and Nikiforov MA (2008) Direct role of nucleotide metabolism in C-MYC-dependent proliferation of melanoma cells. Cell Cycle 7:2392-2400.). Interestingly, c-MYC expression increases purine and pyrimidine metabolism and provides sufficient nucleotide pools to rescue replication stress induced by high rates of DNA synthesis upon disruption of the RB-E2F pathway (Bester et al., 2011Bester AC, Roniger M, Oren YS, Im MM, Sarni D, Chaoat M, Bensimon A, Zamir G, Shewach DS and Kerem B (2011) Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145:435-446.). Furthermore, MYC proteins directly upregulate the expression of certain enzymes involved in DNA replication, such as the WRN helicase (Werner syndrome), a protein involved in the resolution of unusual replication intermediates, and the MRN nuclease (MRE11/RAD50/NBS1), a complex responsible for DSB repair and restart of collapsed replication forks (Grandori et al., 2003Grandori C, Wu KJ, Fernandez P, Ngouenet C, Grim J, Clurman BE, Moser MJ, Oshima J, Russell DW, Swisshelm K et al. (2003) Werner syndrome protein limits MYC-induced cellular senescence. Genes Dev 17:1569-1574.; Robinson et al., 2009Robinson K, Asawachaicharn N, Galloway DA and Grandori C (2009) c-Myc accelerates S-phase and requires WRN to avoid replication stress. PLoS One 4:e5951.; Petroni et al., 2016Petroni M, Sardina F, Heil C, Sahún-Roncero M, Colicchia V, Veschi V, Albini S, Fruci D, Ricci B, Soriani A et al. (2016) The MRN complex is transcriptionally regulated by MYCN during neural cell proliferation to control replication stress. Cell Death Differ 23:197-206.). Upregulation of WRN helicase and MRN nuclease constitute safeguard mechanisms to protect cells from replication stress and DNA damage upon MYC expression.

Oncogenic MYC is frequently associated with human tumorigenesis. As discussed above, MYC overexpression induces replication stress and DSBs, which may be eventually associated with genomic instability. In fact, it has been shown that oncogenic MYC causes chromosomal aberrations, such as deletions, amplifications, and translocations, aneuploidy, and telomeric fusions (Felsher and Bishop, 1999Felsher DW and Bishop JM (1999) Transient excess of MYC activity can elicit genomic instability and tumorigenesis. Proc Natl Acad Sci U S A 96:3940-3944.; Karlsson et al., 2003Karlsson A, Deb-Basu D, Cherry A, Turner S, Ford J and Felsher DW (2003) Defective double-strand DNA break repair and chromosomal translocations by MYC overexpression. Proc Natl Acad Sci U S A 100:9974-9979.; Louis et al., 2005Louis SF, Vermolen BJ, Garini Y, Young IT, Guffei A, Lichtensztejn Z, Kuttler F, Chuang TC, Moshir S, Mougey V et al. (2005) c-Myc induces chromosomal rearrangements through telomere and chromosome remodeling in the interphase nucleus. Proc Natl Acad Sci U S A 102:9613-9618.). Oncogenic MYC has also been shown to induce fragility at specific genomic sites, such as CFS and ERFS (Barlow et al., 2013Barlow JH, Faryabi RB, Callén E, Wong N, Malhowski A, Chen HT, Gutierrez-Cruz G, Sun HW, McKinnon P, Wright G et al. (2013) Identification of early replicating fragile sites that contribute to genome instability. Cell 152:620-632.).

Cyclin E

Cyclin E is one of the prototypical oncogenes that induce replication stress. The Cyclin E family is composed of two proteins, Cyclin E1 and E2 (CCNE1 and CCNE2), which share similar gene sequences and cellular functions. Normally, Cyclin E protein levels peak at the G1/S transition and are completely degraded by the end of S phase. In association with CDK2, Cyclin E controls DNA replication through phosphorylation of multiple proteins, such as the RB tumor suppressor and the DNA replication factors CDT1, CDC6, and Treslin. RB phosphorylation leads to release of E2F transcription factors, which induce the expression of various genes required for DNA replication, while phosphorylation of DNA replication factors is essential for origin licensing and origin firing. Therefore, it is not surprising that oncogenic activation of Cyclin E interferes with cell cycle progression and DNA replication, causing replication stress and genomic instability. CCNE1 amplification, overexpression or impaired protein degradation has been observed in premalignant lesions and cancers, such as breast and lung tumors, and leukemias (Hwang and Clurman, 2005Hwang HC and Clurman BE (2005) Cyclin E in normal and neoplastic cell cycles. Oncogene 24:2776-2786.; Siu et al., 2012Siu KT, Rosner MR and Minella AC (2012) An integrated view of cyclin E function and regulation. Cell Cycle 11:57-64.; Teixeira and Reed, 2017Teixeira LK and Reed SI (2017) Cyclin E deregulation and genomic instability. Adv Exp Med Biol 1042:527-547.).

Many different mechanisms have been shown to contribute to Cyclin E-induced replication stress (Table 1). Unscheduled levels of Cyclin E1 directly interfere with pre-RC formation during late mitosis and early G1 phase, specifically with the recruitment of the helicase subunits MCM2, MCM4, and MCM7 to chromatin (Ekholm-Reed et al., 2004Ekholm-Reed S, Méndez J, Tedesco D, Zetterberg A, Stillman B and Reed SI (2004) Deregulation of cyclin E in human cells interferes with prereplication complex assembly. J Cell Biol 165:789-800.). Inefficient assembly of pre-RC prevents appropriate origin licensing and compromises origin firing and DNA synthesis initiation. Indeed, it has been observed that Cyclin E1 overexpression results in either decreased (Liberal et al., 2012Liberal V, Martinsson-Alzhén HS, Liberal J, Spruck CH, Widschwendter M, McGowan CH and Reed SI (2012) Cyclin-dependent kinase subunit (Cks) 1 or Cks2 overexpression overrides the DNA damage response barrier triggered by activated oncoproteins. Proc Natl Acad Sci U S A 109:2754-2759.) or increased origin firing (Jones et al., 2013Jones RM, Mortusewicz O, Afzal I, Lorvellec M, García P, Helleday T and Petermann E (2013) Increased replication initiation and conflicts with transcription underlie Cyclin E-induced replication stress. Oncogene 32:3744-3753.) in different models. Besides interference with origin licensing and origin firing, Cyclin E overexpression also impairs replication fork progression. High levels of Cyclin E1 cause premature termination of replication forks, fork collapse, and DSBs (Bartkova et al., 2005Bartkova J, Horejsí Z, Koed K, Kramer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C et al. (2005) DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434:864-870., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-). It has been shown that replication fork collapse induced by Cyclin E can be repaired by the homologous recombination pathway break-induced replication (BIR), further leading to copy number alterations and genomic instability (Costantino et al., 2014Costantino L, Sotiriou SK, Rantala JK, Magin S, Mladenov E, Helleday T, Haber JE, Iliakis G, Kallioniemi OP and Halazonetis TD (2014) Break-induced replication repair of damaged forks induces genomic duplications in human cells. Science 343:88-91.). It is important to note that replication stress induced by Cyclin E is dependent on CDK2, as high levels of CDK2 activity are sufficient to impair replication fork progression and cause DNA damage (Hughes et al., 2013Hughes BT, Sidorova J, Swanger J, Monnat RJ Jr and Clurman BE (2013) Essential role for Cdk2 inhibitory phosphorylation during replication stress revealed by a human Cdk2 knockin mutation. Proc Natl Acad Sci U S A 110:8954-8959.).

Another primary mechanism for Cyclin E-induced replication stress is reduction of nucleotide pools. Through disruption of the RB/E2F pathway, Cyclin E1 overexpression enforces cell hyperproliferation with insufficient nucleotide levels, interfering with replication fork progression and causing DSBs (Bester et al., 2011Bester AC, Roniger M, Oren YS, Im MM, Sarni D, Chaoat M, Bensimon A, Zamir G, Shewach DS and Kerem B (2011) Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145:435-446.). Interestingly, cellular supplementation with exogenous nucleosides or induction of nucleotide metabolism through c-MYC expression are able to attenuate replication stress and DNA damage induced by Cyclin E1 overexpression.

Cyclin E-induced replication stress is also caused by transcription-replication collisions, which can lead to DNA topological stress and formation of persistent R-loops. Inhibition of transcription elongation has been shown to alleviate replication stress and reduce DNA damage caused by oncogenic Cyclin E1 (Jones et al., 2013Jones RM, Mortusewicz O, Afzal I, Lorvellec M, García P, Helleday T and Petermann E (2013) Increased replication initiation and conflicts with transcription underlie Cyclin E-induced replication stress. Oncogene 32:3744-3753.). Consistently, inhibition of replication initiation also restores normal levels of fork progression upon high levels of Cyclin E1. Together, these results indicate that oncogenic Cyclin E1 induces replication stress through generation of transcription-replication conflicts. One potential consequence of these encounters is the formation of DNA replication intermediates that are generated in response to topological stress, such as reversed replication forks. Indeed, high levels of Cyclin E1 induce the appearance of aberrant reversed forks (Neelsen et al., 2013Neelsen KJ, Zanini IM, Herrador R and Lopes M (2013) Oncogenes induce genotoxic stress by mitotic processing of unusual replication intermediates. J Cell Biol 200:699-708.).

In a recent work, the human genome has been mapped respective to ORI distribution and replication timing under normal and high levels of Cyclin E1 (Macheret and Halazonetis, 2018Macheret M and Halazonetis TD (2018) Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555:112-116.). Under normal conditions, ORIs are predominantly activated in intergenic regions. Instead, overexpression of Cyclin E1 leads to shortened G1, rapid S phase entry, and novel origin firing in intragenic regions with high transcriptional activity. Excessive origin firing in protein-coding genes facilitates conflicts between transcription and replication machineries, generating replication fork collapse, DSBs, and chromosomal rearrangements (Macheret and Halazonetis, 2018Macheret M and Halazonetis TD (2018) Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555:112-116.).

As discussed before, intrinsic genomic characteristics may sensitize cells to replication stress upon oncogenic insults. In fact, Cyclin E1 deregulation allows cells to enter into mitosis with incomplete replication at specific genomic segments, resulting in mitotic aberrations, such as chromosome breaks and anaphase bridges, as well as CNAs (Teixeira et al., 2015Teixeira LK, Wang X, Li Y, Ekholm-Reed S, Wu X, Wang P and Reed SI (2015) Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 25:1327-1333.). Genomic fragility caused by Cyclin E1 overexpression shows several features of CFS, such as low origin density, late-replicating domains, very long genes, and DNA secondary structures (Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.; Teixeira et al., 2015Teixeira LK, Wang X, Li Y, Ekholm-Reed S, Wu X, Wang P and Reed SI (2015) Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 25:1327-1333.; Teixeira and Reed, 2017Teixeira LK and Reed SI (2017) Cyclin E deregulation and genomic instability. Adv Exp Med Biol 1042:527-547.). Accordingly, genomic breakpoints and rearrangements induced by Cyclin E1 overexpression in vitro are reflected in a large cohort of human cancers with CCNE1 amplification (Zack et al., 2013Zack TI, Schumacher SE, Carter SL, Cherniack AD, Saksena G, Tabak B, Lawrence MS, Zhsng CZ, Wala J, Mermel CH et al. (2013) Pan-cancer patterns of somatic copy number alteration. Nat Genet 45:1134-1140.; Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.; Teixeira et al., 2015Teixeira LK, Wang X, Li Y, Ekholm-Reed S, Wu X, Wang P and Reed SI (2015) Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 25:1327-1333.; Macheret and Halazonetis, 2018Macheret M and Halazonetis TD (2018) Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555:112-116.).

CDC6

CDC6 is a DNA replication-licensing factor that is essential for pre-RC assembly during late mitosis and early G1. Specifically, CDC6 facilitates the loading of MCM helicase to ORIs and is also able to mediate the activation of cell cycle checkpoints and regulate gene transcription (Borlado and Méndez, 2008Borlado LR and Méndez J (2008) CDC6: From DNA replication to cell cycle checkpoints and oncogenesis. Carcinogenesis 29:237-243.). Aberrant expression of CDC6 induces several oncogenic properties in vitro, such as DDR activation, cellular transformation, and genomic instability, as well as tumor growth in vivo. Furthermore, high levels of CDC6 have been observed in advanced stages of non-small cell lung carcinoma (NSCLC) and colon cancer (Bartkova et al., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-; Liontos et al., 2007Liontos M, Koutsami M, Sideridou M, Evangelou K, Kletsas D, Levy B, Kotsinas A, Nahum O, Zoumpourlis V, Kouloukoussa M et al. (2007) Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res 67:10899-10909.; Sideridou et al., 2011Sideridou M, Zakopoulou R, Evangelou K, Liontos M, Kotsinas A, Rampakakis E, Gagos S, Kahata K, Grabusic K, Gkouskou K et al. (2011) Cdc6 expression represses E-cadherin transcription and activates adjacent replication origins. J Cell Biol 195:1123-1140.).

As expected for a protein involved in pre-RC formation, unbalanced levels of CDC6 during cell cycle progression interfere with origin licensing and/or activation (Table 1). The initial evidence for CDC6-induced replication stress came from the observation that overexpression of CDC6, in cooperation with CDT1, promotes origin refiring and DNA rereplication in p53-deficient cells within a few hours of S phase, leading to amplification of large genomic segments and genomic instability (Vaziri et al., 2003Vaziri C, Saxena S, Jeon Y, Lee C, Murata K, Machida Y, Wagle N, Hwang DS and Dutta A (2003) A p53-dependent checkpoint pathway prevents rereplication. Mol Cell 11:997-1008.). Later on, oncogenic CDC6 was confirmed to increase ORI activation at specific genomic sites through chromatin displacement of the CTCF chromosome insulator (Sideridou et al., 2011Sideridou M, Zakopoulou R, Evangelou K, Liontos M, Kotsinas A, Rampakakis E, Gagos S, Kahata K, Grabusic K, Gkouskou K et al. (2011) Cdc6 expression represses E-cadherin transcription and activates adjacent replication origins. J Cell Biol 195:1123-1140.). Additionally, Bartkova and colleagues have shown that high levels of CDC6 induce RPA foci formation, an indicative of ssDNA that has been consistently associated with stalled replication forks (Bartkova et al., 2006Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-).

Besides increased origin firing and DNA rereplication, CDC6-induced replication stress can also occur through collision between replication and transcription machineries and formation of R-loops (Komseli et al., 2018Komseli ES, Pateras IS, Krejsgaard T, Stawiski K, Rizou SV, Polyzos A, Roumelioti FM, Chiourea M, Mourkioti I, Paparouna E et al. (2018) A prototypical non-malignant epithelial model to study genome dynamics and concurrently monitor micro-RNAs and proteins in situ during oncogene-induced senescence. BMC Genomics 19:37.). Interestingly, R-loop formation caused by CDC6 overexpression is preferentially observed within the nucleoli, consistent with the fact that CDC6 is important for transcriptional regulation of the highly repetitive heterochromatic ribosomal DNA (rDNA) (Huang et al., 2016Huang S, Xu X, Wang G, Lu G, Xie W, Tao W, Zhang H, Jiang Q and Zhang C (2016) DNA replication initiator Cdc6 also regulates ribosomal DNA transcription initiation. J Cell Sci 129:1429-1440.). As a result of replication stress, CDC6 upregulation causes a number of structural and numerical chromosome aberrations in different models, with the majority of breakpoints located at CFS (Liontos et al., 2007Liontos M, Koutsami M, Sideridou M, Evangelou K, Kletsas D, Levy B, Kotsinas A, Nahum O, Zoumpourlis V, Kouloukoussa M et al. (2007) Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res 67:10899-10909.; Sideridou et al., 2011Sideridou M, Zakopoulou R, Evangelou K, Liontos M, Kotsinas A, Rampakakis E, Gagos S, Kahata K, Grabusic K, Gkouskou K et al. (2011) Cdc6 expression represses E-cadherin transcription and activates adjacent replication origins. J Cell Biol 195:1123-1140.; Komseli et al., 2018Komseli ES, Pateras IS, Krejsgaard T, Stawiski K, Rizou SV, Polyzos A, Roumelioti FM, Chiourea M, Mourkioti I, Paparouna E et al. (2018) A prototypical non-malignant epithelial model to study genome dynamics and concurrently monitor micro-RNAs and proteins in situ during oncogene-induced senescence. BMC Genomics 19:37.). As discussed before, it is important to consider that CDC6 upregulation may be a consequence of RAS or Cyclin E1 oncogene activation, leading to DNA rereplication, DDR activation, and genomic instability (Mailand and Diffley, 2005Mailand N and Diffley JF (2005) CDKs promote DNA replication origin licensing in human cells by protecting Cdc6 from APC/C-dependent proteolysis. Cell 122:915-926.; Di Micco et al., 2006Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.).

Other oncogenes

Besides the well-characterized roles of RAS, MYC, Cyclin E1, and CDC6 oncoproteins in replication stress, several other oncogenes are also associated with this condition (Table 1). The CDC25 family of proteins is composed of three phosphatases (CDC25A, B, and C) that play critical roles in cell cycle progression and checkpoint control. At particular cell cycle stages and under certain conditions, CDC25 phosphatases directly dephosphorylate and activate CDKs to promote cell cycle transitions. Also, DDR activation triggers CDC25 degradation upon DNA damage, leading to CDK inactivation and cell cycle arrest in order to mediate DNA repair, cell death, or senescence. CDC25 oncogenic properties have been illustrated by cellular transformation, aneuploidy, and tumor formation in vivo, either in cooperation with oncogenic RAS or RB1 loss (Boutros et al., 2007Boutros R, Lobjois V and Ducommun B (2007) CDC25 phosphatases in cancer cells: key players? Good targets? Nat Rev Cancer 7:495-507.). In agreement with its role as an oncogene, CDC25 overexpression has been documented in a variety of human cancers and correlated with disease aggressiveness and poor patient prognosis (Galaktionov et al., 1995Galaktionov K, Lee AK, Eckstein J, Draetta G, Meckler J, Loda M and Beach D (1995) CDC25 phosphatases as potential human oncogenes. Science 269:1575-1577.; Cangi et al., 2000Cangi MG, Cukor B, Soung P, Signoretti S, Moreira G Jr, Ranashinge M, Cady B, Pagano M and Loda M (2000) Role of the Cdc25A phosphatase in human breast cancer. J Clin Invest 106:753-761.).

Initial overexpression of CDC25A causes unscheduled origin activation and DDR induction, while sustained levels of CDC25A leads to checkpoint disruption and chromosomal breaks (Mailand et al., 2000Mailand N, Falck J, Lukas C, Syljuåsen RG, Welcker M, Bartek J and Lukas J (2000) Rapid destruction of human Cdc25A in response to DNA damage. Science 288:1425-1429.; Bartkova et al., 2005Bartkova J, Horejsí Z, Koed K, Kramer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C et al. (2005) DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434:864-870.; Cangi et al., 2008Cangi MG, Piccinin S, Pecciarini L, Talarico A, Dal Cin E, Grassi S, Grizzo A, Maestro R and Doglioni C (2008) Constitutive overexpression of CDC25A in primary human mammary epithelial cells results in both defective DNA damage response and chromosomal breaks at fragile sites. Int J Cancer 123:1466-1471.). Importantly, it has been shown that CDC25A overexpression slows down replication fork progression and induces reversed forks (Neelsen et al., 2013Neelsen KJ, Zanini IM, Herrador R and Lopes M (2013) Oncogenes induce genotoxic stress by mitotic processing of unusual replication intermediates. J Cell Biol 200:699-708.). Besides CDC25A, other members of the CDC25 family also seem to be associated with replication stress, indicating a conserved function for these proteins in regulating cell cycle checkpoints and DDR activation. Increased levels of CDC25B or CDC25C interfere with DNA replication, leading to DNA damage, premature mitotic entry, and chromosomal aberrations (Varmeh and Manfredi, 2009Varmeh S and Manfredi JJ (2009) Inappropriate activation of cyclin-dependent kinases by the phosphatase Cdc25b results in premature mitotic entry and triggers a p53-dependent checkpoint. J Biol Chem 284:9475-9488.; Bugler et al., 2010Bugler B, Schmitt E, Aressy B and Ducommun B (2010) Unscheduled expression of CDC25B in S-phase leads to replicative stress and DNA damage. Mol Cancer 9:29.). However, the molecular mechanisms for these events have not been completely elucidated.

One proto-oncogene that is essential for cell cycle/death control and has been associated with DNA replication stress is the mouse double minute 2 (MDM2) human protein. MDM2 directly interacts with the tumor suppressor p53 to regulate several cellular processes. MDM2 inactivates p53 transactivation domain, promotes its export from the nucleus to the cytoplasm, and induces p53 ubiquitin-mediated degradation. As a negative regulator of p53, it is not surprising that MDM2 amplification and/or overexpression are frequently observed in human cancers, such as many subtypes of sarcomas as well as gliomas and leukemias (Karni-Schmidt et al., 2016Karni-Schmidt O, Lokshin M and Prives C (2016) The roles of MDM2 and MDMX in cancer. Annu Rev Pathol 11:617-644.). It has been shown that MDM2 overexpression inhibits origin firing through activation of the intra S-phase checkpoint, causing unscheduled DNA replication (Frum et al., 2014Frum RA, Singh S, Vaughan C, Mukhopadhyay ND, Grossman SR, Windle B, Deb S and Deb SP (2014) The human oncoprotein MDM2 induces replication stress eliciting early intra-S-phase checkpoint response and inhibition of DNA replication origin firing. Nucleic Acids Res 42:926-940.). Conversely, it has also been shown that p53 activation and subsequent MDM2 upregulation both enhance replication fork progression and increase replication fork processivity (Klusmann et al., 2016Klusmann I, Rodewald S, Müller L, Friedrich M, Wienken M, Li Y, Schulz-Heddergott R and Dobbelstein M (2016) p53 activity results in DNA replication fork processivity. Cell Rep 17:1845-1857.). Although these findings appear conflicting, it is tempting to speculate that disruption of the p53/MDM2 axis in human cancers, either by TP53 mutation or MDM2 overexpression, may interfere with origin firing and replication fork stability. The precise molecular mechanism by which MDM2 overexpression controls origin activation and causes replication stress remains to be determined.

B-cell lymphoma 2 (BCL-2) is another proto-oncogene involved in cell death regulation that has also been linked to replication stress. BCL-2 anti-apoptotic protein promotes cell survival primarily by coordinating protein interactions at several cellular compartments to control mitochondrial membrane permeability. Overexpression of BCL-2 inhibits cell death, facilitates the acquisition of genetic alterations during tumorigenesis, and is frequently observed in human malignancies, including follicular lymphoma, leukemia, and lung carcinoma (Delbridge et al., 2016Delbridge AR, Grabow S, Strasser A and Vaux DL (2016) Thirty years of BCL-2: translating cell death discoveries into novel cancer therapies. Nat Rev Cancer 16:99-109.). Concerning the process of DNA replication, it has been shown that BCL-2 directly inhibits ribonuclease reductase (RNR) activity through binding and disruption of the RRM1/RRM2 complex formation (Xie et al., 2013Xie M, Yen Y, Owonikoko TK, Ramalingam SS, Khuri FR, Curran WJ, Doetsch PW and Deng X (2013) Bcl2 induces DNA replication stress by inhibiting ribonucleotide reductase. Cancer Res 74:212-223.). BCL-2-induced RNR inhibition leads to decreased intracellular levels of dNTPs, slower progression of replication forks, and replication fork asymmetry, all classical features of replication stress.

Oncogenic alterations in the PI3K/AKT signaling pathway represent another insult frequently observed in human cancers. However, alterations in PIK3CA or AKT have not been unequivocally associated with replication stress to date. On the other hand, the PTEN tumor suppressor protein, which counterbalances the PI3K/AKT pathway in the cytoplasm, has been clearly linked to DNA replication, DNA repair, and genome stability in the nucleus (Brandmaier et al., 2017Brandmaier A, Hou SQ and Shen WH (2017) Cell cycle control by PTEN. J Mol Biol 429:2265-2277.; Lee et al., 2018Lee YR, Chen M and Pandolfi PP (2018) The functions and regulation of the PTEN tumour suppressor: new modes and prospects. Nat Rev Mol Cell Biol 19:547-562.). Indeed, it has been shown that PTEN loss impairs replication fork progression and causes replication fork stalling during unperturbed conditions (He et al., 2015He J, Kang X, Yin Y, Chao KS and Shen WH (2015) PTEN regulates DNA replication progression and stalled fork recovery. Nat Commun 6:7620.). Under conditions of replication stress, PTEN is also essential for stability and recovery of stalled replication forks (Feng et al., 2015Feng J, Liang J, Li J, Li Y, Liang H, Zhao X, McNutt MA and Yin Y (2015) PTEN controls the DNA replication process through MCM2 in response to replicative stress. Cell Rep 13:1295-1303.; He et al., 2015He J, Kang X, Yin Y, Chao KS and Shen WH (2015) PTEN regulates DNA replication progression and stalled fork recovery. Nat Commun 6:7620.; Wang et al., 2015Wang G, Li Y, Wang P, Liang H, Cui M, Zhu M, Guo L, Su Q, Sun Y, McNutt MA et al. (2015) PTEN regulates RPA1 and protects DNA replication forks. Cell Res 25:1189-1204.). Several independent mechanisms have been proposed to explain the requirement for PTEN in protecting DNA replication forks. PTEN facilitates the recovery of stalled forks by directly recruiting RAD51 to chromatin, a recombinase that plays multiple roles in DNA replication and repair (He et al., 2015He J, Kang X, Yin Y, Chao KS and Shen WH (2015) PTEN regulates DNA replication progression and stalled fork recovery. Nat Commun 6:7620.). Additionally, upon replication stress, PTEN restricts replication fork progression through dephosphorylation of MCM2, potentially regulating MCM complex function (Feng et al., 2015Feng J, Liang J, Li J, Li Y, Liang H, Zhao X, McNutt MA and Yin Y (2015) PTEN controls the DNA replication process through MCM2 in response to replicative stress. Cell Rep 13:1295-1303.). Finally, PTEN has also been shown to protect replication forks through stabilization of the ssDNA-binding protein RPA1 in a phosphatase-independent manner (Wang et al., 2015Wang G, Li Y, Wang P, Liang H, Cui M, Zhu M, Guo L, Su Q, Sun Y, McNutt MA et al. (2015) PTEN regulates RPA1 and protects DNA replication forks. Cell Res 25:1189-1204.). Together, these studies indicate that PTEN disruption may lead to progressive accumulation of replication errors, DNA damage, and ultimately contribute to genomic instability in cancer.

Conclusions and Perspectives

Normal DNA replication is essential to maintain genome stability in all living organisms. Perturbations in DNA replication may compromise transmission of genetic information to daughter cells, leading to DNA damage and mutations. In fact, increased frequency of DNA replication errors during stem cell divisions has been shown to be associated with higher cancer incidence in humans (Tomasetti and Vogelstein, 2015Tomasetti C and Vogelstein B (2015) Variation in cancer risk among tissues can be explained by the number of stem cell divisions. Science 347:78-81.). In precancerous lesions, one important source of DNA replication errors is oncogene activation, which leads to sustained cellular proliferation and DNA replication stress. Elucidating the causes and consequences of oncogene-induced replication stress is therefore fundamental for better understanding human carcinogenesis.

An extensive body of work has shown that a number of oncogenic insults induce replication stress and genomic instability in human cells. Interestingly, distinct oncogenes, such as H-RAS and CCNE1, are able to generate unique genome fragility landscapes in the same cell type (Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.). As discussed in previous sections, this can be explained by the fact that each oncogene induces replication stress through specific mechanisms. In addition, it is clear that one same replicative insult (either oncogenic or not) causes particular genomic alterations in distinct cell types, including fibroblasts, lymphocytes, and epithelial cells (Le Tallec et al., 2011Le Tallec B, Dutrillaux B, Lachages AM, Millot GA, Brison O and Debatisse M (2011) Molecular profiling of common fragile sites in human fibroblasts. Nat Struct Mol Biol 18:1421-1423., 2013Le Tallec B, Millot GA, Blin ME, Brison O, Dutrillaux B and Debatisse M (2013) Common fragile site profiling in epithelial and erythroid cells reveals that most recurrent cancer deletions lie in fragile sites hosting large genes. Cell Rep 4:420-428.; Hosseini et al., 2013Hosseini SA, Horton S, Saldivar JC, Miuma S, Stampfer MR, Heerema NA and Huebner K (2013) Common chromosome fragile sites in human and murine epithelial cells and FHIT/FRA3B loss-induced global genome instability. Genes Chromosomes Cancer 52:1017-1029.; Miron et al., 2015Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.; Teixeira et al., 2015Teixeira LK, Wang X, Li Y, Ekholm-Reed S, Wu X, Wang P and Reed SI (2015) Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 25:1327-1333.). Specific genomic fragility among different cell types is possibly related to cell-type specific chromatin structure and organization, DNA replication timing, and transcriptional activity among other factors (Alabert and Groth, 2012Alabert C and Groth A (2012) Chromatin replication and epigenome maintenance. Nat Rev Mol Cell Biol 13:153-167.; Sima and Gilbert, 2014Sima J and Gilbert DM (2014) Complex correlations: replication timing and mutational landscapes during cancer and genome evolution. Curr Opin Genet Dev 25:93-100.; Santos-Pereira and Aguilera, 2015Santos-Pereira JM and Aguilera A (2015) R loops: new modulators of genome dynamics and function. Nat Rev Genet 16:583-597.). Together, these observations indicate that replication stress induced by specific oncogenes can create unique repertoires of genomic alterations in different human cell types and cancers.

Replication stress has been considered a potential vulnerability of cancer cells and represents a promising target for cancer therapy. In cancer cells, replication stress may be largely attributed to constitutive oncogene activation. Indeed, multiple signs of oncogene-induced replication stress and consequent DDR pathway activation are frequently observed in precancerous lesions. Recent therapeutic approaches have focused on identifying synthetic lethal interactions between cancer-associated mutations and DNA replication vulnerabilities (Ubhi and Brown, 2019Ubhi T and Brown GW (2019) Exploiting DNA replication stress for cancer treatment. Cancer Res 79:1730-1739.). It has been proposed that, under specific conditions of oncogenic activation, inhibition of DDR proteins induces extensive replication stress, irreversible DSBs, and subsequent cell death, leading to selective elimination of cancer cells. In fact, transformed cells and tumors showing replication stress induced by MYC, RAS, or Cyclin E1 oncoproteins are highly sensitive to ATR or CHK1 kinase inhibitors in different in vitro and in vivo models (Gilad et al., 2010Gilad O, Nabet BY, Ragland RL, Schoppy DW, Smith KD, Durham AC and Brown EJ (2010) Combining ATR suppression with oncogenic Ras synergistically increases genomic instability, causing synthetic lethality or tumorigenesis in a dose-dependent manner. Cancer Res 70:9693-9702.; Murga et al., 2011Murga M, Campaner S, Lopez-Contreras AJ, Toledo LI, Soria R, Montaña MF, Artista L, Schleker T, Guerra C, Garcia E et al. (2011) Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat Struct Mol Biol 18:1331-1335.; Toledo et al., 2011Toledo LI, Murga M, Zur R, Soria R, Rodriguez A, Martinez S, Oyarzabal, J, Pastor J, Bischoff JR and Fernandez-Capetillo O (2011) A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nat Struct Mol Biol 18:721-727.; Schoppy et al., 2012Schoppy DW, Ragland RL, Gilad O, Shastri N, Peters AA, Murga M, Fernandez-Capetillo O, Diehl JA and Brown EJ (2012). Oncogenic stress sensitizes murine cancers to hypomorphic suppression of ATR. J Clin Invest 122:241-252.). Several combined therapies of traditional chemotherapeutic agents with DDR inhibitors are under investigation in clinical trials and have shown promising results to cancer patients. Some of the current challenges for improving the efficacy of replication stress-based therapies consist of identifying particular tumor types that are more likely to respond to specific treatments, determining optimal treatment strategy combinations, and establishing precise therapeutic doses and windows for intervention without generating adverse side effects. Over the coming years, the field of oncogene-induced replication stress will certainly experience further fundamental, exciting discoveries.

Acknowledgments

We apologize to authors whose work has not been cited due to space constraints. This work was supported by grants from The Pew Charitable Trusts, Swiss Bridge, Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and Fundação Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ). LMFP was supported by fellowships from Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) and Ministério da Saúde (INCA).

Conflict of Interest

The authors declare that there is no conflict of interest that could be perceived as prejudicial to the impartiality of the reported research.

Author Contributions

LMFP and LKT contributed equally to the writing of this review.

References

  • Aird KM, Zhang G, Li H, Tu Z, Bitler BG, Garipov A, Wu H, Wei Z, Wagner SN, Herlyn M et al. (2013) Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep 3:1252-1265.
  • Alabert C and Groth A (2012) Chromatin replication and epigenome maintenance. Nat Rev Mol Cell Biol 13:153-167.
  • Aze A, Sannino V, Soffientini P, Bachi A and Costanzo V (2016) Centromeric DNA replication reconstitution reveals DNA loops and ATR checkpoint suppression. Nat Cell Biol 18:684-691.
  • Barlow JH, Faryabi RB, Callén E, Wong N, Malhowski A, Chen HT, Gutierrez-Cruz G, Sun HW, McKinnon P, Wright G et al. (2013) Identification of early replicating fragile sites that contribute to genome instability. Cell 152:620-632.
  • Bartkova J, Horejsí Z, Koed K, Kramer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C et al. (2005) DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434:864-870.
  • Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, Vassiliou LV, Kolettas E, Niforou K, Zoumpourlis VC et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633-637.—-
  • Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, Donovan J, Barretina J, Boehm JS, Dobson J, Urashima M et al. (2010) The landscape of somatic copy-number alteration across human cancers. Nature 463:899-905.
  • Bester AC, Roniger M, Oren YS, Im MM, Sarni D, Chaoat M, Bensimon A, Zamir G, Shewach DS and Kerem B (2011) Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145:435-446.
  • Bignell GR, Greenman CD, Davies H, Butler AP, Edkins S, Andrews JM, Buck G, Chen L, Beare D, Latimer C et al. (2010) Signatures of mutation and selection in the cancer genome. Nature 463:893-898.
  • Black EM and Giunta S (2018) Repetitive fragile sites: Centromere satellite DNA as a source of genome instability in human diseases. Genes 9:E615.
  • Blackford AN and Jackson SP (2017) ATM, ATR, and DNA-PK: The trinity at the heart of the DNA damage response. Mol Cell 66:801-817.
  • Bloom K and Costanzo V (2017) Centromere structure and function. In: Black BE (ed) Centromeres and kinetochores. Springer, Cham, pp 515-539.
  • Bochman ML, Paeschke K and Zakian VA (2012) DNA secondary structures: stability and function of G-quadruplex structures. Nat Rev Genet 13:770-780.
  • Borlado LR and Méndez J (2008) CDC6: From DNA replication to cell cycle checkpoints and oncogenesis. Carcinogenesis 29:237-243.
  • Boutros R, Lobjois V and Ducommun B (2007) CDC25 phosphatases in cancer cells: key players? Good targets? Nat Rev Cancer 7:495-507.
  • Brandmaier A, Hou SQ and Shen WH (2017) Cell cycle control by PTEN. J Mol Biol 429:2265-2277.
  • Bretones G, Delgado MD and León J (2015) Myc and cell cycle control. Biochim Biophys Acta 1849:506-516.
  • Bugler B, Schmitt E, Aressy B and Ducommun B (2010) Unscheduled expression of CDC25B in S-phase leads to replicative stress and DNA damage. Mol Cancer 9:29.
  • Burgers PMJ and Kunkel TA (2017) Eukaryotic DNA replication fork. Annu Rev Biochem 86:417-438.
  • Cangi MG, Cukor B, Soung P, Signoretti S, Moreira G Jr, Ranashinge M, Cady B, Pagano M and Loda M (2000) Role of the Cdc25A phosphatase in human breast cancer. J Clin Invest 106:753-761.
  • Cangi MG, Piccinin S, Pecciarini L, Talarico A, Dal Cin E, Grassi S, Grizzo A, Maestro R and Doglioni C (2008) Constitutive overexpression of CDC25A in primary human mammary epithelial cells results in both defective DNA damage response and chromosomal breaks at fragile sites. Int J Cancer 123:1466-1471.
  • Chambers VS, Marsico G, Boutell JM, Di Antonio M, Smith GP and Balasubramanian S (2015) High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat Biotechnol 33:877-881.
  • Costantino L, Sotiriou SK, Rantala JK, Magin S, Mladenov E, Helleday T, Haber JE, Iliakis G, Kallioniemi OP and Halazonetis TD (2014) Break-induced replication repair of damaged forks induces genomic duplications in human cells. Science 343:88-91.
  • Dang CV (2012) MYC on the path to cancer. Cell 149:22-35.
  • Debatisse M, Le Tallec B, Letessier A, Dutrillaux B and Brison O (2012) Common fragile sites: Mechanisms of instability revisited. Trends Genet 28:22-32.
  • Delbridge AR, Grabow S, Strasser A and Vaux DL (2016) Thirty years of BCL-2: translating cell death discoveries into novel cancer therapies. Nat Rev Cancer 16:99-109.
  • Denko NC, Giaccia AJ, Stringer JR and Stambrook PJ (1994) The human Ha-ras oncogene induces genomic instability in murine fibroblasts within one cell cycle. Proc Natl Acad Sci U S A 91:5124-5128.
  • Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, Schurra C, Garré M, Nuciforo PG, Bensimon A et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638-642.
  • Di Micco R, Fumagalli M and d’Adda di Fagagna F (2007) Breaking news: High-speed race ends in arrest - how oncogenes induce senescence. Trends Cell Biol 17:529-536.
  • Dominguez-Sola D and Gautier J (2014) MYC and the control of DNA replication. Cold Spring Harb Perspect Med 4:a014423.
  • Dominguez-Sola D, Ying CY, Grandori C, Ruggiero L, Chen B, Li M, Galloway DA, Gu W, Gautier J and Dalla-Favera R (2007) Non-transcriptional control of DNA replication by c-Myc. Nature 448:445-451.
  • Durkin SG and Glover TW (2007) Chromosome fragile sites. Annu Rev Genet 41:169-192.
  • Ekholm-Reed S, Méndez J, Tedesco D, Zetterberg A, Stillman B and Reed SI (2004) Deregulation of cyclin E in human cells interferes with prereplication complex assembly. J Cell Biol 165:789-800.
  • Felsher DW and Bishop JM (1999) Transient excess of MYC activity can elicit genomic instability and tumorigenesis. Proc Natl Acad Sci U S A 96:3940-3944.
  • Feng J, Liang J, Li J, Li Y, Liang H, Zhao X, McNutt MA and Yin Y (2015) PTEN controls the DNA replication process through MCM2 in response to replicative stress. Cell Rep 13:1295-1303.
  • Fragkos M, Ganier O, Coulombe P and Méchali M (2015) DNA replication origin activation in space and time. Nat Rev Mol Cell Biol 16:360-374.
  • Frum RA, Singh S, Vaughan C, Mukhopadhyay ND, Grossman SR, Windle B, Deb S and Deb SP (2014) The human oncoprotein MDM2 induces replication stress eliciting early intra-S-phase checkpoint response and inhibition of DNA replication origin firing. Nucleic Acids Res 42:926-940.
  • Gaillard H, García-Muse T and Aguilera A (2015) Replication stress and cancer. Nat Rev Cancer 15:276-289.
  • Galaktionov K, Lee AK, Eckstein J, Draetta G, Meckler J, Loda M and Beach D (1995) CDC25 phosphatases as potential human oncogenes. Science 269:1575-1577.
  • Ge XQ, Jackson DA and Blow JJ (2007) Dormant origins licensed by excess Mcm2-7 are required for human cells to survive replicative stress. Genes Dev 21:3331-3341.
  • Gilad O, Nabet BY, Ragland RL, Schoppy DW, Smith KD, Durham AC and Brown EJ (2010) Combining ATR suppression with oncogenic Ras synergistically increases genomic instability, causing synthetic lethality or tumorigenesis in a dose-dependent manner. Cancer Res 70:9693-9702.
  • Giunta S and Funabiki H (2017) Integrity of the human centromere DNA repeats is protected by CENP-A, CENP-C, and CENP-T. Proc Natl Acad Sci U S A 114:1928-1933.
  • Glover TW, Wilson TE and Arlt MF (2017) Fragile sites in cancer: More than meets the eye. Nat Rev Cancer 17:489-501.
  • Gorgoulis VG, Vassiliou LV, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, Venere M, DiTullio RA Jr, Kastrinakis NG, Levy B et al. (2005) Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434:907-913.
  • Grandori C, Wu KJ, Fernandez P, Ngouenet C, Grim J, Clurman BE, Moser MJ, Oshima J, Russell DW, Swisshelm K et al. (2003) Werner syndrome protein limits MYC-induced cellular senescence. Genes Dev 17:1569-1574.
  • Guerra C, Mijimolle N, Dhawahir A, Dubus P, Barradas M, Serrano M, Campuzano V and Barbacid M (2003) Tumor incidence by and endogenous K-ras oncogene is highly dependent on cellular context. Cancer Cell 4:111-120.
  • Halazonetis TD, Gorgoulis VG and Bartek J (2008) An oncogene-induced DNA damage model for cancer development. Science 319:1352-1355.
  • He J, Kang X, Yin Y, Chao KS and Shen WH (2015) PTEN regulates DNA replication progression and stalled fork recovery. Nat Commun 6:7620.
  • Helmrich A, Ballarino M and Tora L (2011) Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol Cell 44:966-977.
  • Helmrich A, Ballarino M, Nudler E and Tora L (2013) Transcription-replication encounters, consequences and genomic instability. Nat Struct Mol Biol 20:412-418.
  • Higa M, Fujita M and Yoshida K (2017) DNA replication origins and fork progression at mammalian telomeres. Genes 8:E112.
  • Hills SA and Diffley JF (2014) DNA replication and oncogene-induced replicative stress. Curr Biol 24:R435-R444.
  • Hosseini SA, Horton S, Saldivar JC, Miuma S, Stampfer MR, Heerema NA and Huebner K (2013) Common chromosome fragile sites in human and murine epithelial cells and FHIT/FRA3B loss-induced global genome instability. Genes Chromosomes Cancer 52:1017-1029.
  • Huang S, Xu X, Wang G, Lu G, Xie W, Tao W, Zhang H, Jiang Q and Zhang C (2016) DNA replication initiator Cdc6 also regulates ribosomal DNA transcription initiation. J Cell Sci 129:1429-1440.
  • Hughes BT, Sidorova J, Swanger J, Monnat RJ Jr and Clurman BE (2013) Essential role for Cdk2 inhibitory phosphorylation during replication stress revealed by a human Cdk2 knockin mutation. Proc Natl Acad Sci U S A 110:8954-8959.
  • Hussain T, Liu B, Shrock MS, Williams T and Aldaz CM (2019) WWOX, the FRA16D gene: A target of and a contributor to genomic instability. Genes Chromosomes Cancer 58:324-338.
  • Hwang HC and Clurman BE (2005) Cyclin E in normal and neoplastic cell cycles. Oncogene 24:2776-2786.
  • Ibarra A, Schowb E and Méndez J (2008) Excess MCM proteins protect human cells from replicative stress by licensing backup origins of replication. Proc Natl Acad Sci U S A 105:8956-8961.
  • Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, Fearon ER, Sundaresan M, Finkel T and Goldschmidt-Clermont PJ (1997) Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275:1649-1652.
  • Johansson E and Dixon N (2013) Replicative DNA polymerases. Cold Spring Harb Perspect Biol 5:a012799.
  • Jones RM, Mortusewicz O, Afzal I, Lorvellec M, García P, Helleday T and Petermann E (2013) Increased replication initiation and conflicts with transcription underlie Cyclin E-induced replication stress. Oncogene 32:3744-3753.
  • Karlsson A, Deb-Basu D, Cherry A, Turner S, Ford J and Felsher DW (2003) Defective double-strand DNA break repair and chromosomal translocations by MYC overexpression. Proc Natl Acad Sci U S A 100:9974-9979.
  • Karni-Schmidt O, Lokshin M and Prives C (2016) The roles of MDM2 and MDMX in cancer. Annu Rev Pathol 11:617-644.
  • Karnoub AE and Weinberg RA (2008) Ras oncogenes: Split personalities. Nat Rev Mol Cell Biol 9:517-531.
  • Kaushal S and Freudenreich CH (2019) The role of fork stalling and DNA structures in causing chromosome fragility. Genes Chromosomes Cancer 58:270-283.
  • Klusmann I, Rodewald S, Müller L, Friedrich M, Wienken M, Li Y, Schulz-Heddergott R and Dobbelstein M (2016) p53 activity results in DNA replication fork processivity. Cell Rep 17:1845-1857.
  • Komseli ES, Pateras IS, Krejsgaard T, Stawiski K, Rizou SV, Polyzos A, Roumelioti FM, Chiourea M, Mourkioti I, Paparouna E et al. (2018) A prototypical non-malignant epithelial model to study genome dynamics and concurrently monitor micro-RNAs and proteins in situ during oncogene-induced senescence. BMC Genomics 19:37.
  • Kotsantis P, Silva LM, Irmscher S, Jones RM, Folkes L, Gromak N and Petermann E (2016) Increased global transcription activity as a mechanism of replication stress in cancer. Nat Commun 7:13087.
  • Kotsantis P, Petermann E and Boulton SJ (2018) Mechanisms of oncogene-induced replication stress: Jigsaw falling into place. Cancer Discov 8:537-555.
  • Lane AN and Fan TW (2015) Regulation of mammalian nucleotide metabolism and biosynthesis. Nucleic Acids Res 43:2466-2485.
  • Le Beau MM, Rassool FV, Neilly ME, Espinosa R 3rd, Glover TW, Smith DI and McKeithan TW (1998) Replication of a common fragile site, FRA3B, occurs late in S phase and is delayed further upon induction: implications for the mechanism of fragile site induction. Hum Mol Genet 7:755-761.
  • Le Tallec B, Dutrillaux B, Lachages AM, Millot GA, Brison O and Debatisse M (2011) Molecular profiling of common fragile sites in human fibroblasts. Nat Struct Mol Biol 18:1421-1423.
  • Le Tallec B, Millot GA, Blin ME, Brison O, Dutrillaux B and Debatisse M (2013) Common fragile site profiling in epithelial and erythroid cells reveals that most recurrent cancer deletions lie in fragile sites hosting large genes. Cell Rep 4:420-428.
  • Lee AC, Fenster BE, Ito H, Takeda K, Bae NS, Hirai T, Yu ZX, Ferrans VJ, Howard BH and Finkel T (1999) Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J Biol Chem 274:7936-7940.
  • Lee YR, Chen M and Pandolfi PP (2018) The functions and regulation of the PTEN tumour suppressor: new modes and prospects. Nat Rev Mol Cell Biol 19:547-562.
  • Leikam C, Hufnagel A, Schartl M and Meierjohann S (2008) Oncogene activation in melanocytes links reactive oxygen to multinucleated phenotype and senescence. Oncogene 27:7070-7082.
  • Leonard AC and Méchali M (2013) DNA replication origins. Cold Spring Harb Perspect Biol 5:a010116.
  • Letessier A, Millot GA, Koundrioukoff S, Lachagès AM, Vogt N, Hansen RS, Malfoy B, Brison O and Debatisse M (2011) Cell-type-specific replication initiation programs set fragility of the FRA3B fragile site. Nature 470:120-123.
  • Liberal V, Martinsson-Alzhén HS, Liberal J, Spruck CH, Widschwendter M, McGowan CH and Reed SI (2012) Cyclin-dependent kinase subunit (Cks) 1 or Cks2 overexpression overrides the DNA damage response barrier triggered by activated oncoproteins. Proc Natl Acad Sci U S A 109:2754-2759.
  • Liontos M, Koutsami M, Sideridou M, Evangelou K, Kletsas D, Levy B, Kotsinas A, Nahum O, Zoumpourlis V, Kouloukoussa M et al. (2007) Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res 67:10899-10909.
  • Liu YC, Li F, Handler J, Huang CR, Xiang Y, Neretti N, Sedivy JM, Zeller KI and Dang CV (2008) Global regulation of nucleotide biosynthetic genes by c-Myc. PLoS One 3:e2722.
  • Louis SF, Vermolen BJ, Garini Y, Young IT, Guffei A, Lichtensztejn Z, Kuttler F, Chuang TC, Moshir S, Mougey V et al. (2005) c-Myc induces chromosomal rearrangements through telomere and chromosome remodeling in the interphase nucleus. Proc Natl Acad Sci U S A 102:9613-9618.
  • Lujan SA, Williams JS and Kunkel TA (2016) DNA polymerases divide the labor of genome replication. Trends Cell Biol 26:640-654.
  • Macheret M and Halazonetis TD (2015) DNA replication stress as a hallmark of cancer. Annu Rev Pathol 10:425-448.
  • Macheret M and Halazonetis TD (2018) Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555:112-116.
  • Machida YJ, Teer JK and Dutta A (2005) Acute reduction of an origin recognition complex (ORC) subunit in human cells reveals a requirement of ORC for Cdk2 activation. J Biol Chem 280:27624-27630.
  • Mailand N and Diffley JF (2005) CDKs promote DNA replication origin licensing in human cells by protecting Cdc6 from APC/C-dependent proteolysis. Cell 122:915-926.
  • Mailand N, Falck J, Lukas C, Syljuåsen RG, Welcker M, Bartek J and Lukas J (2000) Rapid destruction of human Cdc25A in response to DNA damage. Science 288:1425-1429.
  • Mallette FA, Gaumont-Leclerc MF and Ferbeyre G (2007) The DNA damage signaling pathway is a critical mediator of oncogene-induced senescence. Genes Dev 21:43-48.
  • Mannava S, Grachtchouk V, Wheeler LJ, Im M, Zhuang D, Slavina EG, Mathews CK, Shewach DS and Nikiforov MA (2008) Direct role of nucleotide metabolism in C-MYC-dependent proliferation of melanoma cells. Cell Cycle 7:2392-2400.
  • Martínez P and Blasco MA (2015) Replicating through telomeres: A means to an end. Trends Biochem Sci 40:504-515.
  • Martínez P, Thanasoula M, Muñoz P, Liao C, Tejera A, McNees C, Flores JM, Fernández-Capetillo O, Tarsounas M and Blasco MA (2009) Increased telomere fragility and fusions resulting from TRF1 deficiency lead to degenerative pathologies and increased cancer in mice. Genes Dev 23:2060-2075.
  • Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N and Oda M (2010) Eukaryotic chromosome DNA replication: where, when, and how? Annu Rev Biochem 79:89-130.
  • Maya-Mendoza A, Ostrakova J, Kosar M, Hall A, Duskova P, Mistrik M, Merchut-Maya JM, Hodny Z, Bartkova J, Christensen C et al. (2015) Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol Oncol 9:601-616.
  • McIntosh D and Blow JJ (2012) Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb Perspect Biol 4:a012955.
  • Méchali M (2010) Eukaryotic DNA replication origins: Many choices for appropriate answers. Nat Rev Mol Cell Biol 11:728-738.
  • Miron K, Golan-Lev T, Dvir R, Ben-David E and Kerem B (2015) Oncogenes create a unique landscape of fragile sites. Nat Commun 6:7094.
  • Mortusewicz O, Herr P and Helleday T (2013) Early replication fragile sites: Where replication-transcription collisions cause genetic instability. EMBO J 32:493-495.
  • Murga M, Campaner S, Lopez-Contreras AJ, Toledo LI, Soria R, Montaña MF, Artista L, Schleker T, Guerra C, Garcia E et al. (2011) Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat Struct Mol Biol 18:1331-1335.
  • Neelsen KJ and Lopes M (2015) Replication fork reversal in eukaryotes: from dead end to dynamic response. Nat Rev Mol Cell Biol 16:207-220.
  • Neelsen KJ, Zanini IM, Herrador R and Lopes M (2013) Oncogenes induce genotoxic stress by mitotic processing of unusual replication intermediates. J Cell Biol 200:699-708.
  • O’Donnell M, Langston L and Stillman B (2013) Principles and concepts of DNA replication in bacteria, archea, and eukarya. Cold Spring Harb Perspect Biol 5:a010108.
  • Ozeri-Galai E, Lebofsky R, Rahat A, Bester AC, Bensimon A and Kerem B (2011) Failure of origin activation in response to fork stalling leads to chromosomal instability at fragile sites. Mol Cell 43:122-131.
  • Ozeri-Galai E, Bester AC and Kerem B (2012) The complex basis underlying common fragile site instability in cancer. Trends Genet 28:295-302.
  • Petroni M, Sardina F, Heil C, Sahún-Roncero M, Colicchia V, Veschi V, Albini S, Fruci D, Ricci B, Soriani A et al. (2016) The MRN complex is transcriptionally regulated by MYCN during neural cell proliferation to control replication stress. Cell Death Differ 23:197-206.
  • Pylayeva-Gupta Y, Grabocka E and Bar-Sagi D (2011) RAS oncogenes: Weaving a tumorigenic web. Nat Rev Cancer 11:761-774.
  • Rai P, Young JJ, Burton DG, Giribaldi MG, Onder TT and Weinberg RA (2011) Enhanced elimination of oxidized guanine nucleotides inhibits oncogenic RAS-induced DNA damage and premature senescence. Oncogene 30:1489-1496.
  • Ray S, Atkuri KR, Deb-Basu D, Adler AS, Chang HY, Herzenberg LA and Felsher DW (2006) MYC can induce DNA breaks in vivo and in vitro independent of reactive oxygen species. Cancer Res 66:6598-6605.
  • Robinson K, Asawachaicharn N, Galloway DA and Grandori C (2009) c-Myc accelerates S-phase and requires WRN to avoid replication stress. PLoS One 4:e5951.
  • Rohban S and Campaner S (2015) Myc induced replicative stress response: How to cope with it and exploit it. Biochim Biophys Acta 1849:517-254.
  • Saldivar JC and Park D (2019) Mechanisms shaping the mutational landscape of the FRA3B/FHIT-deficient cancer genome. Genes Chromosomes Cancer 58:317-323.
  • Saldivar JC, Cortez D and Cimprich KA (2017) The essential kinase ATR: Ensuring faithful duplication of a challenging genome. Nat Rev Mol Cell Biol 18:622-636.
  • Sankar N, Kadeppagari RK and Thimmapaya B (2009) c-Myc-induced aberrant DNA synthesis and activation of DNA damage response in p300 knockdown cells. J Biol Chem 284:15193-15205.
  • Santos-Pereira JM and Aguilera A (2015) R loops: new modulators of genome dynamics and function. Nat Rev Genet 16:583-597.
  • Schoppy DW, Ragland RL, Gilad O, Shastri N, Peters AA, Murga M, Fernandez-Capetillo O, Diehl JA and Brown EJ (2012). Oncogenic stress sensitizes murine cancers to hypomorphic suppression of ATR. J Clin Invest 122:241-252.
  • Sfeir A, Kosiyatrakul ST, Hockemeyer D, MacRae SL, Karlseder J, Schildkraut CL and de Lange T (2009) Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138:90-103.
  • Shreeram S, Sparks A, Lane DP and Blow JJ (2002) Cell type-specific responses of human cells to inhibition of replication licensing. Oncogene 21:6624-6632.
  • Siddiqui K, On KF and Diffley JF (2013) Regulating DNA replication in eukarya. Cold Spring Harb Perspect Biol 5:a012930.
  • Sideridou M, Zakopoulou R, Evangelou K, Liontos M, Kotsinas A, Rampakakis E, Gagos S, Kahata K, Grabusic K, Gkouskou K et al. (2011) Cdc6 expression represses E-cadherin transcription and activates adjacent replication origins. J Cell Biol 195:1123-1140.
  • Sima J and Gilbert DM (2014) Complex correlations: replication timing and mutational landscapes during cancer and genome evolution. Curr Opin Genet Dev 25:93-100.
  • Sirbu BM and Cortez D (2013) DNA damage response: Three levels of DNA repair regulation. Cold Spring Harb Perspect Biol 5:a012724.
  • Siu KT, Rosner MR and Minella AC (2012) An integrated view of cyclin E function and regulation. Cell Cycle 11:57-64.
  • Srinivasan SV, Dominguez-Sola D, Wang LC, Hyrien O and Gautier J (2013) Cdc45 is a critical effector of Myc-dependent DNA replication stress. Cell Rep 3:1629-1639.
  • Suram A, Kaplunov J, Patel PL, Ruan H, Cerutti A, Boccardi V, Fumagalli M, Di Micco R, Mirani N, Gurung RL et al. (2012) Oncogene-induced telomere dysfunction enforces cellular senescence in human cancer precursor lesions. EMBO J 31:2839-2851.
  • Tanaka S and Araki H (2013) Helicase activation and establishment of replication forks at chromosomal origins of replication. Cold Spring Harb Perspect Biol 5:a010371.
  • Técher H, Koundrioukoff S, Nicolas A and Debatisse M (2017) The impact of replication stress on replication dynamics and DNA damage in vertebrate cells. Nat Rev Genet 18:535-550.
  • Teixeira LK and Reed SI (2017) Cyclin E deregulation and genomic instability. Adv Exp Med Biol 1042:527-547.
  • Teixeira LK, Wang X, Li Y, Ekholm-Reed S, Wu X, Wang P and Reed SI (2015) Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 25:1327-1333.
  • Toledo LI, Murga M, Zur R, Soria R, Rodriguez A, Martinez S, Oyarzabal, J, Pastor J, Bischoff JR and Fernandez-Capetillo O (2011) A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nat Struct Mol Biol 18:721-727.
  • Toledo L, Neelsen KJ and Lukas J (2017) Replication catastrophe: when a checkpoint fails because of exhaustion. Mol Cell 66:735-749.
  • Tomasetti C and Vogelstein B (2015) Variation in cancer risk among tissues can be explained by the number of stem cell divisions. Science 347:78-81.
  • Tsantoulis PK, Kotsinas A, Sfikakis PP, Evangelou K, Sideridou M, Levy B, Mo L, Kittas C, Wu XR, Papavassiliou AG et al. (2008) Oncogene-induced replication stress preferentially targets common fragile sites in preneoplastic lesions. A genome-wide study. Oncogene 27:3256-3264.
  • Tu Z, Aird KM, Bitler BG, Nicodemus JP, Beeharry N, Xia B, Yen TJ and Zhang R (2011) Oncogenic RAS regulates BRIP1 expression to induce dissociation of BRCA1 from chromatin, inhibit DNA repair, and promote senescence. Dev Cell 21:1077-1091.
  • Tubbs A, Sridharan S, van Wietmarschen N, Maman Y, Callen E, Stanlie A, Wu W, Wu X, Day A, Wong N et al. (2018) Dual roles of poly(dA:dT) tracts in replication initiation and fork collapse. Cell 174:1127-1142.
  • Ubhi T and Brown GW (2019) Exploiting DNA replication stress for cancer treatment. Cancer Res 79:1730-1739.
  • Vafa O, Wade M, Kern S, Beeche M, Pandita TK, Hampton GM and Wahl GM (2002) c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability. Mol Cell 9:1031-1044.
  • Varmeh S and Manfredi JJ (2009) Inappropriate activation of cyclin-dependent kinases by the phosphatase Cdc25b results in premature mitotic entry and triggers a p53-dependent checkpoint. J Biol Chem 284:9475-9488.
  • Vaziri C, Saxena S, Jeon Y, Lee C, Murata K, Machida Y, Wagle N, Hwang DS and Dutta A (2003) A p53-dependent checkpoint pathway prevents rereplication. Mol Cell 11:997-1008.
  • Wang G, Li Y, Wang P, Liang H, Cui M, Zhu M, Guo L, Su Q, Sun Y, McNutt MA et al. (2015) PTEN regulates RPA1 and protects DNA replication forks. Cell Res 25:1189-1204.
  • Weyemi U, Lagente-Chevallier O, Boufraqech M, Prenois F, Courtin F, Caillou B, Talbot M, Dardalhon M, Al Ghuzlan A, Bidart JM et al. (2012) ROS-generating NADPH oxidase NOX4 is a critical mediator in oncogenic H-Ras-induced DNA damage and subsequent senescence. Oncogene 31:1117-1129.
  • Xie M, Yen Y, Owonikoko TK, Ramalingam SS, Khuri FR, Curran WJ, Doetsch PW and Deng X (2013) Bcl2 induces DNA replication stress by inhibiting ribonucleotide reductase. Cancer Res 74:212-223.
  • Zack TI, Schumacher SE, Carter SL, Cherniack AD, Saksena G, Tabak B, Lawrence MS, Zhsng CZ, Wala J, Mermel CH et al. (2013) Pan-cancer patterns of somatic copy number alteration. Nat Genet 45:1134-1140.
  • Zeman MK and Cimprich KA (2014) Causes and consequences of replication stress. Nat Cell Biol 16:2-9.
  • Associate Editor: Carlos F. M. Menck

Publication Dates

  • Publication in this collection
    13 Dec 2019
  • Date of issue
    2020

History

  • Received
    23 Apr 2019
  • Accepted
    09 July 2019
Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
E-mail: editor@gmb.org.br