Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Oct 15.
Published in final edited form as: FEBS Lett. 2019 Oct 15;593(20):2853–2867. doi: 10.1002/1873-3468.13619

Preparation for DNA Replication: The Key to a Successful S phase

Juanita C Limas 1, Jeanette Gowen Cook 1,2
PMCID: PMC6817399  NIHMSID: NIHMS1052297  PMID: 31556113

Abstract

Successful genome duplication is required for cell proliferation and demands extraordinary precision and accuracy. The mechanisms by which cells enter, progress through, and exit S phase are intense areas of focus in the cell cycle and genome stability fields. Key molecular events in the G1 phase of the cell division cycle, especially origin licensing, are essential for pre-establishing conditions for efficient DNA replication during the subsequent S phase. If G1 events are poorly regulated or disordered, then DNA replication can be compromised leading to genome instability, a hallmark of tumorigenesis. Upon entry into S phase, coordinated origin firing and replication progression ensure complete, timely, and precise chromosome replication. Both G1 and S phase progression are controlled by master cell cycle protein kinases and ubiquitin ligases that govern the activity and abundance of DNA replication factors. In this short review, we describe current understanding and recent developments related to G1 progression and S phase entrance and exit with a particular focus on origin licensing regulation in vertebrates.

Keywords: cell cycle, DNA replication, origin licensing, genome stability, replication stress, checkpoint

Introduction

The somatic eukaryotic cell cycle is a series of phases: G1, S, G2, and mitosis. Cells start the cell cycle in G1 (gap phase 1) progress through S (DNA replication), G2 (gap phase 2), and then divide in M (mitosis) [1]. Cells enter G1 either from the preceding mitosis and cytokinesis or from a quiescent state (also known as G0 phase) which is distinct from the replicative cell cycle. Cells transit between G0 and the replicative cycle in response to a variety of extracellular mitogenic signals [2, 3].

Genome stability, the faithful duplication and transmission of chromosomal DNA from one cell to its daughter cells, depends on accurate and efficient DNA replication. Events that prepare for DNA replication must be ordered and precise to avoid catastrophic consequences and genome instability which includes increased frequencies of base pair mutations, duplication or deletion of chromosomal loci, variation in chromosome number and structure, and changes in microsatellites (tandem nucleotide repeats scattered throughout the genome) [4]. Molecular events in G1 set the stage for DNA replication in S phase. Likewise, a collection of molecular processes governs orderly progression through S phase in preparation for G2 phase. Decades of intense investigation using a variety of experimental systems have identified key DNA replication proteins and their regulators in G1 and S. Nevertheless, important questions remain: How exactly do cells “know” when to transition from G1 into S phase

Origin licensing prepares cells for S phase by loading inactive replicative helicases onto DNA during G1. To achieve complete but precise genome duplication, licensing must start after mitosis and finish prior to the initiation of S phase when origin firing takes place. Origin firing involves helicase activation in S phase and establishing bidirectional replication forks, but once origins have fired, no new licensing is permitted [5, 6] (Figure 1A). The consequences of any new origin licensing in S phase are serious because re-replication creates extra copies of loci that can lead to genome instability (Figure 1B). On the other hand, too few licensed origins can also lead to genome instability associated with incomplete replication (Figure 1C).

Figure 1. Illustrations of normal and abnormal origin licensing outcomes.

Figure 1.

A. Normal origin licensing and CDK2-dependent firing in G1 and S phase. B. Unscheduled licensing and firing leading to re-replication and genome instability. C. Underlicensing in G1 leading to replication stress and genome instability.

G1: Preparing DNA for S phase

A primary activity during G1 is preparing DNA for replication during S phase. Replication initiates at specific sites known as DNA replication origins, and efficient eukaryotic replication requires many active origins per chromosome [7, 8]. What defines an origin varies by organism: among eukaryotes some yeasts such as S. cerevisiae have origins that are (at least partly) defined by conserved DNA sequences [912]. While origins have been successfully mapped in other eukaryotes, their locations are not easily predicted from DNA sequences. Highly active replication origins in several vertebrates have recently been shown to have some common general features such as open chromatin, GC-rich sequences, G4 quadruplex structures, or proximity to active genes (reviewed in [12, 13]).

Certain histone modifications play an active role in regulating the chromatin landscapes relevant to origin licensing. In particular, monomethylation of histone H4 lysine 20 (H4K20me1) promotes origin licensing. The monomethyltransferase SETD8 (also known as PR-Set7) catalyzes H4K20me1. Depleting SETD8 causes defects in origin licensing, whereas inappropriate SETD8 expression in S phase promotes unregulated re-licensing and re-replication[1416]. Subsequent tri-methylation of histone H4K20, which requires H4K20me1 as a substrate, ensures that late firing origins are activated during S phase. Furthermore, replication origins are active even in heterochromatin where lower accessibility likely makes licensing and origin activation challenging. Some proteins that are not required at origins in other genomic regions, such as ORCA [17] and shelterin [18], are important for licensing heterochromatin or telomeres. Considering the complexity of the eukaryotic genome, it is remarkable that enough origins are licensed in G1 to ensure the entire genome is usually replicated in S phase with few problems.

Origin licensing involves the sequential action of licensing factors to load the core of the replicative helicase, MCM, onto DNA (Figure 2, text box, and Table 1). Our understanding of the licensing mechanism has advanced significantly from biochemical reconstitutions using S. cerevisiae or X. laevis licensing proteins. Origin licensing starts with ORC (Origin Recognition Complex, a DNA binding complex of six subunits Orc1–6), which directly binds origins of replication. ORC works with two additional helicase loading factors, CDC6 (Cell Division Cycle 6) and CDT1 (CDC10-dependent transcript 1), to load at least two hexamers of the MCM replicative helicase (Mini-chromosome Maintenance Complex, six subunits MCM2–7) [19, 20]. Licensed origins have at least one loaded MCM double hexamer (= two MCM hexamers). MCM complexes form the core of the replicative helicases at replication forks but are not active during G1; they will be activated or “fired” in S phase. Once MCM double hexamers are loaded, the loading factors ORC, CDC6, and CDT1 are not required for MCM complexes to remain on DNA [21]. MCM complexes are stably DNA-bound (reviewed in [20]).

Figure 2. Origin licensing.

Figure 2.

The concerted action of ORC, CDC6, and CDT1 load hexamers of MCM onto DNA during G1 phase. A loaded MCM double hexamer constitutes a licensed origin; See also the text box.

Table.

Glossary of Primary Terms: Vertebrate proteins and complexes in this mini-review* Active Cell Cycle Phase
Abbreviation Full Name Function G1 S G2-M
APC/C Anaphase promoting complex/Cyclosome E3 ubiquitin ligase, targets many substrates from anaphase through G1; inactive in S and G2 (✔)
ATM Ataxia Telangiectasia Mutated protein kinase activated by DNA damage, blocks S phase and M phase entry and stimulates DNA repair
ATR ATM-related protein kinase active during replication stress, and by DNA damage (all phases), arrests/delays cell cycle progression and stimulates DNA repair
CDC45 Cell Division Cycle 45 component of the CMG helicase
CDC6 Cell Division Cycle 6 loads MCM, degraded in early G1, stabilized in late G1, cytoplasmic in S/G2
CDC20 Cell Division Cycle 20 substrate coactivator for the APC/C ubiquitin E3 ligase in M phase
CDH1 CDC20 Homolog 1 substrate coactivator for the APC/C ubiquitin E3 ligase in G1 phase
CDK Cyclin Dependent Kinase protein kinase heterodimer of one cyclin with one catalytic subunit, different CDK in different phases
CDT1 Cdc10-Dependent Transcript 1 loads MCM2–7, degraded in S
CDT2 Cdc10 Dependent Transcript 2 substrate adaptor for the CRL4CDT2 E3 ubiquitin ligase, activated by substrate binding to DNA-loaded PCNA
CHK1 Checkpoint Kinase 1 protein kinase activated by ATR after replication stress or DNA damage
CHK2 Checkpoint Kinase 2 protein kinase activated by ATM after DNA damage
CMG Cdc45 + MCM2–7 + GINS active form of the replicative DNA helicase
CRL4 Cullin Ring Ligase 4 E3 ubiquitin ligase complex, recruited by CDT2 to target CDT1, SETD8, and p21 in S phase
cyclin A Cyclin A subunit of cyclin A/CDK2 and cyclin A/CDK1 kinase
cyclin D Cyclin D activates CDK4 and CDK6 to phosphorylate Rb and contribute to Rb inactivation
cyclin E Cyclin E subunit of cyclin E/CDK2 kinase, stabilizes CDC6, stimulates origin firing
DBF4 Dumbbell Forming Protein 4 subunit of DBF4/CDC7 kinase (DDK)
DDK DBF4 dependent kinase protein kinase heterodimer of DBF4 and CDC7; phosphorylates MCM4 & MCM6
E2F Adenovirus E2 gene Factor transcriptional regulator of genes encoding many key S phase proteins
EMI1 Early mitotic inhibitor 1 inhibits the APC/C ubiquitin E3 ligase during S & G2 phases
GINS Go-Ichi-Ni-San Heterotetramer component of the CMG helicase (SLD5, PSF1–3)
GMNN Geminin inhibits re-replication by blocking CDT1-MCM binding
MCM Mini-chromosome Maintenance heterohexamer of subunits MCM2–7, loaded at origins in G1, component of CMG helicase
ORC Origin Recognition Complex heterohexamer of subunits ORC1–6, binds DNA and loads MCM
p21 protein 21 kDa inhibitor of CDK2 and CDK1
p27 protein 27 kDa inhibitor of CDK2 and CDK1
p53 protein 53 kDa transcription factor induced by DNA damage and other stresses
PCNA Proliferating Cell Nuclear Antigen DNA polymerase clamp/processivity factor at replication forks; coordinates multiple interactions
Rb Retinoblastoma protein transcriptional repressor of E2F-regulated genes; phosphorylated by cyclin D/CDK4, cyclin D/CDK6, and cyclin E/CDK2
SCF SKP1-CUL1-F-box E3 ubiquitin ligase complex, targets some CDK-phosphorylated substrates
SETD8 Su(var)3–9, enhancer-of-zeste, and trithorax histone h4 lysine 20 monomethylase, promotes licensing (a.k.a. PR-SET7, SET8)
SKP2 Suppressor of Kinetochore Protein Mutant 2 F-box protein family member, substrate adapter for SCFSKP2
*

Additional firing and replication fork factors: TOPBP1, treslin/MTBP, MCM10, RECQL4, POLε, RPA, RFC

Origin licensing is promoted in G1 through a combination of transcriptional activation of the genes encoding licensing proteins and post-transcriptional control. CDC6, CDT1, and each subunit of ORC and MCM are the products of genes controlled by the E2F transcription factor family [22, 23]. In G1 phase, mitogen stimulation triggers production of cyclin D which binds and activates the CDK4 and CDK6 protein kinases. Cyclin D/CDK4–6 complexes de-repress E2F by mono-phosphorylating its inhibitor, Rb, at one of many individual phosphorylation sites [24]. Recent work has shown that the 14 principal sites of Rb mono-phosphorylation lead to different profiles of E2F-dependent transcription, indicating considerable complexity in the G1 transcriptional program [25]. In late G1 phase one of the many E2F-activated genes, CCNE1 which encodes the cyclin E protein, is sharply activated [24]. Cyclin E binds and activates the CDK2 protein kinase (cyclin-dependent protein kinase 2) which then hyper-phosphorylates and inactivates Rb in a positive feedback loop. As a result, late G1 cells synthesize considerably more of the licensing proteins than early G1 cells.

At least one of the origin licensing proteins, CDC6, also requires post-transcriptional activation during G1 phase. CDC6 is a substrate of the APC/CCDH1 E3 ubiquitin ligase which is active from mitosis through G1. As a result, CDC6 protein levels remain low during G1 because it is ubiquitylated by APC/CCDH1 and degraded by the proteasome. In late G1 phase however, CDC6 is no longer degraded because APC/CCDH1 cannot bind and ubiquitylate CDC6 once it has been phosphorylated by cyclin E/CDK2 [26]. CDC6 is only one of the 14 polypeptides that form the core of the licensing system (MCM2–7, ORC1–6, CDT1, and CDC6), and each of these proteins is subject to a variety of post-translational modifications at some point in the cell cycle. Many of these modifications are associated with inactivating licensing outside of G1 phase (i.e. in the previous S, G2, and M phases), and thus must be reversed to activate licensing during G1. For example, CDT1 is hyperphosphorylated in mitosis, and those phosphorylation events are associated with inactivating CDT1 for licensing [27]. Similarly, ORC subunits bind chromatin poorly in mitosis as a result of mitotic CDK activity [28, 29]. Presumably these inactivating phosphorylations are reversed by phosphatases in G1, or the substrate proteins are degraded and resynthesized. In support of phosphatases as activators of licensing, the Rif1-PP1 phosphatase has recently been shown to protect the largest ORC subunit from premature phosphorylation and degradation during G1 phase [30].

Origin licensing checkpoint.

Origin licensing is restricted to G1 phase to avoid re-replication (described below). When has enough origin licensing taken place during G1 to ensure complete genome duplication? In other words, is there a mechanism that ensures sufficient genome-wide origin licensing prior to the activation of S phase? Such a mechanism would be an “origin licensing checkpoint,” and it would delay S phase entry while origin licensing is incomplete. It is clear that in budding yeast no such mechanism exists because null mutations in essential origin licensing proteins do not block cell cycle progression; such mutants enter mitosis with un-replicated DNA [31, 32]. On the other hand, evidence that a licensing checkpoint in mammalian cells delays the activation of the S phase kinase CDK2 when origin licensing is inhibited is described below.

By what criteria can we infer the existence of a mammalian origin licensing checkpoint? In 1988, Hartwell described cell cycle checkpoints as “control mechanisms enforcing dependency in the cell cycle.” Checkpoints [33] refer to signaling pathways that ensure proper completion of key cell cycle steps before the next cell cycle phase transition, but checkpoint regulators are not themselves essential to the processes they regulate. This important aspect of checkpoints is sometimes only evident when one identifies treatments or mutations that bypass the regulation to allow premature cell cycle progression. An analogy for a checkpoint is the roller coaster in Figure 3. Before the ride begins, safety monitors conduct a series of safety checks of the riders themselves and of the equipment. These monitors will not allow the ride to begin unless the rules have been satisfied. If the monitors are absent, then a ride could, in principle, proceed, albeit with added risk. Cell cycle checkpoints are evident when a component of the monitoring system is eliminated that then allows inappropriate or risky cell cycle progression.

Figure. 3. Progression through G1 and S phase represented as a roller-coaster track.

Figure. 3.

Major decision points in G1 and at the G1/S transition are indicated in yellow, and a cell cycle checkpoint at or just prior to the G1/S transition is marked in green.

Evidence for an origin licensing checkpoint is not simply less DNA replication when MCM loading is reduced. A true licensing checkpoint would operate during G1 phase and prevent initial S phase entry. Moreover, the checkpoint regulators must not themselves be essential DNA replication proteins, but rather proteins that restrain S phase entry (i.e. safety monitors). When the checkpoint regulators are depleted or mutated, then cells can proceed prematurely into S phase, albeit with a higher risk of incomplete replication. Indeed, in non-transformed mammalian cell lines, manipulations that inhibit origin licensing result in a longer G1 phase and low CDK2 activity [3336]. Interestingly, the same manipulations in transformed cell lines do not affect G1 length or CDK2 activity, and cells instead proceed into S phase with considerably less than their typical amount of licensed origins. The subsequent S phase is therefore “underlicensed” (Figure 1C). In the most severe licensing depletions, cells experience replication failure and apoptosis [35, 37]. Perhaps the most compelling evidence for a true licensing checkpoint is the fact that removing p53, a central player in the response to cellular stresses that is absent or compromised in tumor derived cells, bypassed all of the checkpoint behaviors. Like transformed cells, non-transformed cells lacking p53 do not delay in G1 with low CDK2 activity even when licensing is impaired [34, 35], and, like their yeast orthologs, can even enter premature mitosis [38]. Thus, it is clear that p53-proficient cells can couple the activation of CDK2 and S phase with the status of origin licensing.

This checkpoint has long been under-investigated because it cannot be studied in either budding or fission yeast (there is no yeast p53 ortholog) or the more conveniently manipulated tumor cell lines. Multiple studies have documented the origin licensing checkpoint phenomenon [36, 3945], though the molecular details of this pathway are still very poorly understood. For example, it is not known if p53 is activated by post-translational modification in the same ways it is activated during a DNA damage response or if basal p53 activity is sufficient to confer licensing checkpoint proficiency. There is little evidence that p53 accumulates after licensing inhibition, suggesting that the strong stabilizing signals that characterize responses to strong DNA damage signals are not part of the licensing checkpoint. Some licensing checkpoint experiments have detected increased p21 protein, and since p21 is the product of a p53-inducible gene, the increased p21 could be an indicator of p53 activation. On the other hand, high p21 is also an indirect consequence of G1 arrest, since p21 is degraded in S phase cells and stable in G1 [46]. We recently found that cells re-entering the cell cycle from quiescence have an impaired licensing checkpoint [34]. This natural checkpoint deficiency may provide a unique avenue to dissect the pathway connecting licensing to S phase entry. Considerable future work will need to determine what exactly cells are measuring about MCM loading, how that measurement is made, how the information is converted into CDK2 regulation, and the role of p53 in that process.

In contrast to the licensing checkpoint, a very well-studied checkpoint is the p53-dependent G1/S checkpoint. This network of surveillance proteins is activated by DNA damage to block S phase entry. DNA damage in G1 triggers ATM and/or ATR kinases depending on the nature of the damage. These kinases stimulate signaling cascades that block cell cycle progression and activate DNA repair (reviewed in [47, 48]). The DNA damage checkpoint does not prevent DNA from being damaged but rather responds to damage and blocks cells from progressing into S phase by inducing inhibition of CDK2 and allowing time for repair.

S phase entry

The large size of eukaryotic genomes presents challenges for timely and efficient DNA replication. Meeting this challenge requires that cells prepare many origins in G1 and activate many origins in S phase. To start S phase, replication origins are activated or “fired” through a series of sequential protein phosphorylation events. Interestingly, not all origins are fired at the same time but rather, are fired at different times as S phase proceeds. Heterochromatin loci can pose more of a challenge for origin firing and thus, are generally replicated later in S phase than the more accessible euchromatic regions. Presumably origins in euchromatin are activated before origins in heterochromatin because the heterochromatic landscape restricts access to firing factors [7].

The replication timing program ensures the temporal order of origin firing (early versus late origin firing) and may be influenced by origin licensing parameters that were established in the preceding G1 phase. High-resolution protein-DNA interaction mapping experiments in budding yeast detected more MCM loading at early-firing origins compared to late-firing origins. This extra MCM could compete for limiting concentrations of firing factors during S phase [49]. Similarly, more ORC has been detected at early versus late-replicating regions in human cells [50]. Interestingly however, this differential loading may not have translated to more origin licensing because no differences in MCM loading were detected in late-replicating heterochromatin versus early replicating euchromatin in human cells [8]. Thus, MCM loading differences do not fully account for establishing replication timing programs. Additional proteins such as the ORC-associated ORCA protein, which binds different origins at different times during G1 [51], may distinguish early versus late origins. Such proteins could either preferentially delay late origins or preferentially accelerate early origins, and such factors could be preloaded in G1 or act exclusively in S phase. Furthermore, topologically associated domains (TADs), three-dimensional chromosome conformations comprised of self-interacting regions of the genome, have also been suggested to influence replication timing. Discrete cis-regulatory units have recently been shown to shape the chromatin landscape and architecture to ensure properly timed replication [52].

Origin firing creates the replisome, or replication “machine”: a protein complex that is assembled at replication origins that were licensed in G1 (reviewed in [5356]). The replisome is responsible for DNA replication in a bidirectional fashion from each origin that fires. Replication forks initiate where DNA is unwound by the replicative helicase. Helicase activation requires CDK2 with contributions from DDK (DBF4-dependent kinase). The replicative helicase is assembled from the MCM complexes that were loaded in G1 during origin licensing to which the CDC45 protein and GINS complex are added during origin firing. This assembly is the CMG (CDC45-MCM-GINS) helicase (reviewed in [53, 54]). GINS and CDC45 are recruited to MCM through TopBP1 and CDK-phosphorylated treslin, a TopBP1-binding protein homologous to the yeast Sld3 initiation protein. In budding yeast, phosphorylation of residues on the MCM2–7 helicase by DDK triggers helicase activation by recruiting Sld3 to mediates CDC45 loading [54, 56]. By extension, vertebrate DDK may function similarly and is required for S phase completion [55]. Origin firing also depends on RECQL4, the MCM10 protein, and DNA polymerase (Pol ɛ). Once the helicase is activated, the bifurcation of two opposing replication forks from the origin commences. Finally, PCNA, RPA, and RFC, among other proteins, are recruited to fully activate the replisome. These proteins are the products of E2F-regulated genes, and thus their availability in S phase requires their prior accumulation in G1 as part of the preparation process [22].

Since origin firing is dependent on both CDK and DDK activity, the activation of these protein kinases in S phase is essential for genome duplication. For DDK, activity depends on accumulation of the DBF4 subunit, a product of an E2F-regulated gene, which binds to the CDC7 catalytic subunit to form active DDK [22]. In the case of CDK2, full activation requires multiple molecular events. First, cyclin protein must accumulate because it is an essential activator of CDK2. In late G1 and early S phase, the cyclin most responsible for CDK2 activation is cyclin E which, as discussed above, is produced by E2F-dependent activation of the CCNE gene. CDK2 must also be phosphorylated at the activating T160 by CDK-activating kinase (CAK) and also lack phosphorylation at Wee1-dependent inhibitory sites T14 and Y15 [57].

Two primary CDK2 inhibitor proteins, p21 and p27, are abundant in G1 and must be degraded in late G1 for full CDK2 activation. The p27 CDK inhibitor is ubiquitylated by the SCFSKP2 E3 ubiquitin ligase after p27 has been phosphorylated by CDK2 [58]. Thus, CDK2 stimulates its own activation by triggering the destruction of one of its inhibitors. The p21 CDK2 inhibitor is also an SCFSKP2 substrate, and SCF targeting causes p21 levels to fall beginning in late G1. The largest drop in p21 levels occurs in early S phase and is dependent on the CRL4CDT2 E3 ubiquitin ligase. Targeting by CRL4CDT2 requires recruitment of both the E3 ubiquitin ligase and the p21 substrate to DNA-loaded PCNA at replication forks in a process termed replication-coupled destruction [46, 5961]. The precise dynamics of CDK inhibitor degradation is a combination of the rate of ubiquitylation and the action of deubiquitinases that oppose ubiquitylation [62]. To date, the role of deubiquitinases that govern p21 and p27 degradation at G1/S are not fully known, though a recent study indicates an important contribution of USP11 for p21 control [63] or USP37 for p27 control [64].

Dynamics at G1/S.

Once a cell passes through the G1/S transition and has commenced origin firing, it is committed to full genome duplication and ultimately, cell division. This commitment is irreversible which necessitates that many molecular events switch robustly and in a coordinated fashion. Our understanding of precise inter-relationships at the G1/S transition has been advanced in recent years by single cell studies using live-cell imaging combined with fluorescent reporters of specific molecular events [6568].

In addition to the many factors contributing to CDK2 activation described thus far, a master E3 ubiquitin ligase is integrated into the G1/S commitment. The anaphase promoting complex, also called the cyclosome or APC/C, is an E3 ubiquitin ligase that functions in G1 to keep mitotic CDKs inactive by targeting S phase and mitotic cyclins for destruction. APC/C also targets multiple proteins directly or indirectly involved in origin firing and activation, including CDC6 (discussed above), the DBF4 subunit of the DDK kinase, the SKP2 component of the SCFSKP2 ubiquitin ligase, and the re-replication inhibitor protein Geminin (discussed below). Robust S phase entry requires that APC/C be active in G1 but inactive during S phase [66, 69, 70].

APC/C activity is dependent on one of two substrate coactivator subunits: CDC20 in M phase and CDH1 in G1 phase. In G1 phase, neither cyclins D nor E are substrates of APC/C, whereas the S phase and M phase cyclins A and B are actively degraded by APC/C. As CDK2 activity rises, CDK2 phosphorylates CDH1 to contribute to initial APC/C inactivation since phosphorylated CDH1 cannot bind APC/C well. CDH1 is subsequently degraded in S phase by SCFcyclin F-dependent ubiquitination[70], an additional contribution to APC/C inactivation [68]. In addition, E2F-stimulated transcription drives production of the APC/C inhibitor protein EMI1 whose levels rise through the G1/S transition (reviewed in [71, 72]). EMI1 is subsequently stabilized by a variety of ubiquitination and phosphorylation events near the G1/S transition [73]. Once APC/C is inactivated, DNA replication (i.e. S phase) is destined to begin. Recent work has underscored APC/C inactivation as an indicator of passage through this “point of no return” regardless of stressors that could have blocked entry into S phase had they been encountered in G1 (e.g. DNA damage) [66]. Whether early S phase can start before APC/C is fully inactive however, is not known.

During S phase

As cells transition into S phase, CDK2 is activated first by cyclin E (early S phase) and then by cyclin A (mid-to-late S phase) to fire origins. Cyclin A accumulation is the result of both E2F-dependent transcription during preparation for S phase in G1 and the early S phase inactivation of APC/C which ubiquitylates cyclin A from early mitosis through the end of G1 phase. CDK2 activity rises from the overall increase in cyclin concentration and the full destruction of both p21 and p27 CDK inhibitors. In mid-to-late S phase, cyclin E is downregulated by a ubiquitylation pathway that, similar the CDK inhibitors, is stimulated by substrate phosphorylation. In the case of cyclin E, the ubiquitin E3 ligase is SCFFbw7 which targets phosphorylated cyclin E [74, 75]. By this point, cyclin A levels are high meaning that during S phase, CDK2 switches binding partners from cyclin E to cyclin A. Although not all CDK2 substrates can be phosphorylated by both cyclin E/CDK2 and cyclin A/CDK2, both forms of CDK2 share the substrates that are phosphorylated during origin firing. In mid-to-late S phase, cyclin A/CDK2 is thus responsible for origin firing later in S phase to ensure replication completion [76, 77].

Re-replication.

As mentioned earlier, origins are not fired at the same time in S phase. Throughout S phase, there are regions of the genome that have already been replicated and regions that remain un-replicated. Genome stability requires that no region be replicated more than once. Thus, mechanisms have evolved to ensure that no sequences become re-replicated in a single cell cycle. The consequences of re-replication are serious and include DNA breaks, replication stress, gene amplification, mutagenesis, chromosome instability, and oncogenesis [7882]. Re-replication is blocked by preventing origin licensing once S phase starts ensuring that even the earliest regions to be replicated are not re-licensed. Vertebrate genomes are large and complex, and S phase is typically many hours, so no single molecular mechanism can prevent all re-licensing. Instead, multiple mechanisms operate in parallel, and all of them are required for full re-replication control [78, 81]. Even seemingly minor perturbations that increase the likelihood of re-replication have serious organismal consequences from the increased genome instability [80, 81, 83].

In vertebrate cells, early S phase is marked by the destruction of both the CDT1 MCM loading protein and the SETD8 histone monomethytransferase. Both of these proteins are targeted by CRL4CDT2 as soon as origins fire because the trigger for CRL4CDT2 targeting is the presence of active PCNA-containing replication forks [59]. CDT1 overproduction or failure to degrade either CDT1 or SETD8 can induce substantial re-replication [16, 84]. During S phase (once APC/C is inactive), the geminin protein is stabilized because it is no longer targeted by APC/C. Geminin binds tightly to CDT1 and blocks interaction with MCM, thus inhibiting CDT1’s licensing function [8587]. This inhibition may be particularly important in G2 phase when CDT1 becomes stable again and re-accumulates both to perform an essential mitotic function and in preparation for the next G1 phase [88, 89]. Even minimal licensing in G2 would, by definition, occur on DNA that had been replicated already, and the high CDK activity in G2 cells could rapidly stimulate re-firing and re-replication at such sites [90, 91].

In S phase, cyclin A/CDK2 specifically inactivates some licensing proteins that are apparently poor substrates for cyclin E/CDK2 (and thus not inactivated in late G1 by cyclin E/CDK2). This specificity between cyclin E and cyclin A preserves licensing activity through the end of G1, but as cyclin A accumulates it inhibits licensing during S and G2 phases. Cyclin A/CDK2 phosphorylates a number of licensing proteins during S phase: (1) CDT1, creating a phosphodegron recognized by the SCFSKP2 E3 ubiquitin ligase to contribute to CDT1 degradation in S phase [92, 93]; (2) CDC6 (near its nuclear localization signal), promoting CDC6 export to the cytoplasm [94]; and (3) ORC1 (the largest subunit of ORC), stimulating Orc1 degradation via SCFSKP2 [94, 95]. At least one additional subunit of ORC, ORC2, is also phosphorylated by CDK2 and/or CDK1 which is activated by cyclin A in G2 and by cyclin B in mitosis. ORC2 phosphorylation reduces its affinity for chromatin [28, 29].

S phase progression is also under control of the intra-S phase checkpoint that responds to DNA damage, to slowed or stalled replication, and to the occasional replication stress that occurs even during a normal, unperturbed S phase. Various types of DNA damage activate the ATR and ATM pathways, which phosphorylate CHK1 and CHK2 kinases respectively. One of the downstream consequences of checkpoint activation is S and M phase CDK inhibition, which slows or stops the pace of new origin firing and blocks mitotic entry (reviewed in [9698]). As a consequence of turning off the origin firing kinases CDK2 and DDK, the intra-S phase checkpoint blocks origin firing, avoiding replication of damaged DNA, protecting replication forks, and also stimulating DNA repair [48, 91, 97, 99]. Even in the absence of DNA damage, ATR controls the pace of origin firing [99, 100].

Interestingly, the intra-S checkpoint does not suppress all origin firing. In the absence of DNA damage, origins fire in clusters as indicated by single-molecule in vivo DNA labelling experiments that detect origin firing [101, 102]. The ATR-CHK1 pathway suppresses origin firing in clusters that have not yet initiated, but licensed origins within an active cluster are not suppressed. It is not yet known how the intra-S checkpoint pathway distinguishes origins that should be suppressed from those that should not.

Dormant origins.

DNA fiber analysis has been used to estimate the replicon size in somatic mammalian cells relative to the number of licensed origins per nucleus, and the number of licensed origins greatly exceeds the number that fire [8, 103]. The number of licensed origins can be reduced substantially by partially depleting MCM subunits or MCM loading factors with little impact on short-term overall proliferation [104, 105]. Why do cells license extra origins during G1 that typically remain dormant during S? Origin efficiency, the likelihood that a particular origin will fire in a given S phase, varies by location and can be influenced by cell type and by local gene expression [106]. It may be that a universal replication licensing program is established in G1, but that cell type or environmental signals that impact local chromatin structure restrict origin firing to a selected subset.

The strongest explanation for why cells license many dormant origins comes from analyzing cellular responses to replication stress and long-term impacts on genome stability [107]. Some replication forks stall because they encounter lesions and blocks during S phase. A suite of lesion bypass, replication fork restart, and DNA repair systems are available during S phase largely because the genes encoding these proteins are also E2F-regulated and therefore induced during G1 in preparation for S phase [47]. If a replication fork stalls and cannot be repaired, then the DNA ahead of the stalled fork must be replicated by a fork converging from the opposite direction. Licensing more origins in G1 ensures cells are prepared to fire these “backup” or dormant origins in S phase in the locations where problems arise. Much like a savings account stores money for unexpected expenses, dormant origins serve as a “savings account” for the genome. Dormant origins (extra origins licensed in G1) are activated or fired in S phase in response to problems encountered during replication [104, 108]. These dormant origins escape suppression by the intra-S checkpoint to generate rescue replication forks near a stalled replisome [81, 104, 105, 108, 109]. If dormant origins are not available, the consequences could be dire (Figure 1C): incomplete replication of the genome could lead to genome instability, hypersensitivity to replication stress, and possibly trigger tumorigenesis if left unchecked [110].

S phase exit

DNA replication should complete before the onset of mitosis. As described above, origin firing during S phase is regulated by the ATR-CHK1 pathway, which simultaneously prevents premature mitosis [100]. The ATR pathway is activated by ongoing replication and blocks not only mitosis itself, but also the transcriptional program that stimulates mitotic protein accumulation; this relationship has been proposed as an S-G2 checkpoint [111]. In addition to ATR, cyclin A/CDK1 is also important for regulating late S phase origin firing [76]. Because cyclin A is an APC/C substrate, and APC/C is inactive during S and G2 phase, cyclin A accumulates to high levels in late S phase to activate both CDK2 and CDK1. Premature activation of cyclin A/CDK1 caused aberrant replication earlier in S phase as a result of dysregulated origin firing. Furthermore, inhibiting CDK1 lengthened the time cells spent in S phase, as cells struggled with inefficient late origin firing at the end of S phase [76]. By these measurements, Cyclin A/CDK1 has a relatively unique role in controlling late S phase events, and high cyclin A/CDK1 activity is one of the few molecular markers of G2 phase.

Essentially all of the MCM that was originally loaded during origin licensing in G1 phase is unloaded by the end of S phase [112]. Most of the dormant origins will not fire, and presumably these MCM complexes are removed by the replication forks that transit those sites. What happens when a replication fork encounters a loaded and unfired MCM double hexamer is currently unknown. MCM complexes in active CMG complexes at replication forks are removed during replisome disassembly. Two proposed mechanisms have been reported: 1) association with an alternate Mcm2 subunit, MCM-BP, has been suggested to unload MCM2–7 from DNA during S phase [113, 114] (though MCM-BP has many partners and proposed roles), 2) polyubiquitylation of the Mcm7 subunit stimulates Mcm7 removal from the complex by a non-degradative mechanism promoting complete replisome disassembly [115117] and reviewed in [118120]. However, many questions remain about the mechanisms and order of events of MCM unloading at the end of S phase.

Conclusion

In summary, the events of G1 play critical roles that determine how well cells are prepared for DNA replication in the subsequent S phase. The sequence of molecular events in G1 and in S phase are currently being further elucidated by extensive use of quantitative high-content single-cell analysis and single-molecule studies. Generating and analyzing time-lapse images of cells with fluorescent cell cycle reporters has proven quite useful for gathering robust data to update our understanding of cell cycle dynamics at both the G1/S and S/G2 transitions [65, 67, 68]. We anticipate that such endeavors will continue to address the relationships among the critical pathways in G1 and S phase. The kinetics of origin licensing during G1 relative to E2F-depedent transcription, CDK activity, and APC/C inactivation are not yet known for normal vertebrate cells. Moreover, how genetic perturbations associated with oncogenic transformation impact licensing kinetics and the downstream consequences of such perturbations in S phase are also incompletely understood. In addition, the precise relationships among epigenetic regulators, licensing locations, and subsequent origin firing by CDK and DDK are still unclear and could be better understood with these newer techniques. In addition, we still don’t fully understand the relationship between replication licensing and the G1/S transition nor the relationship between replication completion and the onset of G2. Because genome stability depends on the efficiency of DNA replication, further studies of how cells enter and exit S phase are still crucial.

Text Box: Origin Licensing is the process of loading MCM hexamers onto DNA by ORC, Cdc6, and Cdt1; see Figure 2.

The molecular details of loading MCM complexes to form DNA-bound double hexamers have been largely determined with purified S. cerevisiae (budding yeast) proteins in vitro. MCM loading in other species is presumed to follow a similar sequence.

1:

  • ORC binds DNA directly

  • Cdc6 binds ORC

  • Cdt1 binds a single MCM2–7 hexamer

  • The Cdt1-MCM2–7 complex binds ORC-Cdc6

  • ORC-Cdc6-Cdt1 load the MCM2–7 hexamer onto DNA; MCM encircles dsDNA

  • MCM hydrolyzes ATP, and Cdc6 and Cdt1 are released

2:

  • Cdc6 binds ORC

  • A second Cdt1-MCM2–7 complex interacts with the loaded ORC-Cdc6-MCM. Alternatively, a second ORC may bind DNA on the opposite side of the loaded MCM.

  • The second MCM2–7 is loaded with the two N-termini of the MCM complexes interacting

  • ORC, Cdc6, and Cdt1 are released, leaving a loaded MCM double hexamer on ds DNA

Acknowledgements.

We are grateful to colleagues for helpful discussions and feedback, and we appreciate the understanding of our colleagues whose work we could not cite because of space restrictions. The authors are supported by grants from the National Institutes of Health (R01GM083024, R01GM102413, T32GM007040, and R25GM055336), and by an HHMI Gilliam Fellowship for Advanced Study to J.C.L (GT10886).

References

  • 1.Morgan DO (2007) The Cell Cycle: Principles of Control, OUP/New Science Press. [Google Scholar]
  • 2.O’Farrell PH (2011) Quiescence: early evolutionary origins and universality do not imply uniformity. Philos Trans R Soc Lond B Biol Sci. 366, 3498–3507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Terzi MY, Izmirli M & Gogebakan B (2016) The cell fate: senescence or quiescence. Mol Biol Rep. 43, 1213–1220. [DOI] [PubMed] [Google Scholar]
  • 4.Aguilera A & Garcia-Muse T (2013) Causes of genome instability. Annu Rev Genet. 47, 1–33. [DOI] [PubMed] [Google Scholar]
  • 5.Riera A, Tognetti S & Speck C (2014) Helicase loading: how to build a MCM2–7 double-hexamer. Semin Cell Dev Biol. 30, 104–109. [DOI] [PubMed] [Google Scholar]
  • 6.Siddiqui K, On KF & Diffley JF (2013) Regulating DNA replication in eukarya. Cold Spring Harb Perspect Biol. 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Rivera-Mulia JC & Gilbert DM (2016) Replicating Large Genomes: Divide and Conquer. Mol Cell. 62, 756–765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Sugimoto N, Maehara K, Yoshida K, Ohkawa Y & Fujita M (2018) Genome-wide analysis of the spatiotemporal regulation of firing and dormant replication origins in human cells. Nucleic Acids Res. 46, 6683–6696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Li H & Stillman B (2012) The origin recognition complex: a biochemical and structural view. Subcell Biochem. 62, 37–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Liachko I, Youngblood RA, Tsui K, Bubb KL, Queitsch C, Raghuraman MK, Nislow C, Brewer BJ & Dunham MJ (2014) GC-rich DNA elements enable replication origin activity in the methylotrophic yeast Pichia pastoris. PLoS Genet. 10, e1004169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Peng C, Luo H, Zhang X & Gao F (2015) Recent advances in the genome-wide study of DNA replication origins in yeast. Front Microbiol. 6, 117–123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Prioleau M-N & MacAlpine DM (2016) DNA replication origins—where do we begin? Genes & Development. 30, 1683–1697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ganier O, Prorok P, Akerman I & Mechali M (2019) Metazoan DNA replication origins. Curr Opin Cell Biol. 58, 134–141. [DOI] [PubMed] [Google Scholar]
  • 14.Brustel J, Kirstein N, Izard F, Grimaud C, Prorok P, Cayrou C, Schotta G, Abdelsamie AF, Dejardin J, Mechali M, Baldacci G, Sardet C, Cadoret JC, Schepers A & Julien E (2017) Histone H4K20 tri-methylation at late-firing origins ensures timely heterochromatin replication. EMBO J. 36, 2726–2741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Smith OK & Aladjem MI (2014) Chromatin structure and replication origins: determinants of chromosome replication and nuclear organization. J Mol Biol. 426, 3330–3341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Tardat M, Brustel J, Kirsh O, Lefevbre C, Callanan M, Sardet C & Julien E (2010) The histone H4 Lys 20 methyltransferase PR-Set7 regulates replication origins in mammalian cells. Nat Cell Biol. 12, 1086–1093. [DOI] [PubMed] [Google Scholar]
  • 17.Giri S, Aggarwal V, Pontis J, Shen Z, Chakraborty A, Khan A, Mizzen C, Prasanth KV, Ait-Si-Ali S, Ha T & Prasanth SG (2015) The preRC protein ORCA organizes heterochromatin by assembling histone H3 lysine 9 methyltransferases on chromatin. Elife. 4, 1–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Ogawa S, Kido S, Handa T, Ogawa H, Asakawa H, Takahashi TS, Nakagawa T, Hiraoka Y & Masukata H (2018) Shelterin promotes tethering of late replication origins to telomeres for replication-timing control. EMBO J. 37, 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Gambus A, Khoudoli GA, Jones RC & Blow JJ (2011) MCM2–7 form double hexamers at licensed origins in Xenopus egg extract. J Biol Chem. 286, 11855–11864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Remus D & Diffley JF (2009) Eukaryotic DNA replication control: lock and load, then fire. Curr Opin Cell Biol. 21, 771–777. [DOI] [PubMed] [Google Scholar]
  • 21.Bowers JL, Randell JC, Chen S & Bell SP (2004) ATP hydrolysis by ORC catalyzes reiterative Mcm2–7 assembly at a defined origin of replication. Mol Cell. 16, 967–978. [DOI] [PubMed] [Google Scholar]
  • 22.Grant GD, Brooks L 3rd, Zhang X, Mahoney JM, Martyanov V, Wood TA, Sherlock G, Cheng C & Whitfield ML (2013) Identification of cell cycle-regulated genes periodically expressed in U2OS cells and their regulation by FOXM1 and E2F transcription factors. Mol Biol Cell. 24, 3634–3650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Leone G, DeGregori J, Yan Z, Jakoi L, Ishida S, Williams RS & Nevins JR (1998) E2F3 activity is regulated during the cell cycle and is required for the induction of S phase. Genes & Development. 12, 2120–2130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Narasimha AM, Kaulich M, Shapiro GS, Choi YJ, Sicinski P & Dowdy SF (2014) Cyclin D activates the Rb tumor suppressor by mono-phosphorylation. Elife. 3, 1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Sanidas I, Morris R, Fella KA, Rumde PH, Boukhali M, Tai EC, Ting DT, Lawrence MS, Haas W & Dyson NJ (2019) A Code of Mono-phosphorylation Modulates the Function of RB. Mol Cell. 73, 985–1000 e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Mailand N & Diffley JF (2005) CDKs promote DNA replication origin licensing in human cells by protecting Cdc6 from APC/C-dependent proteolysis. Cell. 122, 915–26. [DOI] [PubMed] [Google Scholar]
  • 27.Chandrasekaran S, Tan TX, Hall JR & Cook JG (2011) Stress-stimulated mitogen-activated protein kinases control the stability and activity of the Cdt1 DNA replication licensing factor. Mol Cell Biol. 31, 4405–4416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Lee KY, Bang SW, Yoon SW, Lee SH, Yoon JB & Hwang DS (2012) Phosphorylation of ORC2 protein dissociates origin recognition complex from chromatin and replication origins. J Biol Chem. 287, 11891–11898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Li CJ, Vassilev A & DePamphilis ML (2004) Role for Cdk1 (Cdc2)/cyclin A in preventing the mammalian origin recognition complex’s largest subunit (Orc1) from binding to chromatin during mitosis. Mol Cell Biol. 24, 5875–5886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Hiraga SI, Ly T, Garzon J, Horejsi Z, Ohkubo YN, Endo A, Obuse C, Boulton SJ, Lamond AI & Donaldson AD (2017) Human RIF1 and protein phosphatase 1 stimulate DNA replication origin licensing but suppress origin activation. EMBO Rep. 18, 403–419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Hofmann JFX & Beach D (1994) cdt 1 is an essential target of the Cdc1 O/Sct 1 transcription factor: requirement for DNA replication and inhibition of mitosis. EMBO J. 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Piatti S, Lengauerl C & Nasmyth K (1995) Cdc6 is an unstable protein whose de novo synthesis in G1 is important for the onset of S phase and for preventing a ‘reductional’ anaphase in the budding yeast Saccharomyces cerevisiae. EMBO J. 14, 3788–3799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hartwell LH & Weinert TA (1989) Checkpoints: Controls That Ensure the Order of Cell Cycle Events. Science. 246, 629–634. [DOI] [PubMed] [Google Scholar]
  • 34.Matson JP, House AM, Grant GD, Wu H, Perez J & Cook JG (2019) Intrinsic checkpoint deficiency during cell cycle re-entry from quiescence. J Cell Biol. 218, 2169–2184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Nevis KR, Cordeiro-Stone M & Cook JG (2009) Origin licensing and p53 status regulate Cdk2 activity during G(1). Cell Cycle. 8, 1952–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Sotillo E, Garriga J, Padgaonkar A, Kurimchak A, Cook JG & Grana X (2009) Coordinated activation of the origin licensing factor CDC6 and CDK2 in resting human fibroblasts expressing SV40 small T antigen and cyclin E. J Biol Chem. 284, 14126–14135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Shreeram S, Sparks A, Lane DP & Blow JJ (2002) Cell type-specific responses of human cells to inhibition of replication licensing. Oncogene. 21, 6624–6632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lemmens B, Hegarat N, Akopyan K, Sala-Gaston J, Bartek J, Hochegger H & Lindqvist A (2018) DNA Replication Determines Timing of Mitosis by Restricting CDK1 and PLK1 Activation. Mol Cell. 71, 117–128 e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Zimmerman KM, Jones RM, Petermann E & Jeggo PA (2013) Diminished Origin-Licensing Capacity Specifically Sensitizes Tumor Cells to Replication Stress. Mol Cancer Res. 11, 370–380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Gardner NJ, Gillespie PJ, Carrington JT, Frearson JA, Woodland A & Blow JJ (2017) The High-Affinity Interaction between ORC and DNA that Is Required for Replication Licensing Is Inhibited by 2-Arylquinolin-4-Amines. Cell Chemical Biology. 24, 981–992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Matson JP, Dumitru R, Coryell P, Baxley RM, Chen W, Twaroski K, Webber BR, Tolar J, Bielinsky AK, Purvis JE & Cook JG (2017) Rapid DNA replication origin licensing protects stem cell pluripotency. Elife. 6, 1–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Machida YJ, Teer JK & 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. [DOI] [PubMed] [Google Scholar]
  • 43.Teer JK, Machida YJ, Labit H, Novac O, Hyrien O, Marheineke K, Zannis-Hadjopoulos M & Dutta A (2006) Proliferating human cells hypomorphic for origin recognition complex 2 and pre-replicative complex formation have a defect in p53 activation and Cdk2 kinase activation. J Biol Chem. 281, 6253–6260. [DOI] [PubMed] [Google Scholar]
  • 44.Liu P, Slater DM, Lenburg M, Nevis K, Cook JG & Vaziri C (2009) Replication licensing promotes cyclin D1 expression and G1 progression in untransformed human cells. Cell Cycle. 8, 125–136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lunn CL, Chrivia JC & Baldassare JJ (2010) Activation of Cdk2/Cyclin E complexes is dependent on the origin of replication licensing factor Cdc6 in mammalian cells. Cell Cycle. 9, 4533–4541. [DOI] [PubMed] [Google Scholar]
  • 46.Coleman KE, Grant GD, Haggerty RA, Brantley K, Shibata E, Workman BD, Dutta A, Varma D, Purvis JE & Cook JG (2015) Sequential replication-coupled destruction at G1/S ensures genome stability. Genes Dev. 29, 1734–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hustedt N & Durocher D (2016) The control of DNA repair by the cell cycle. Nat Cell Biol. 19, 1–9. [DOI] [PubMed] [Google Scholar]
  • 48.Sancar A, Lindsey-Boltz LA, Unsal-Kacmaz K & Linn S (2004) Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem. 73, 39–85. [DOI] [PubMed] [Google Scholar]
  • 49.Das SP, Borrman T, Liu VW, Yang SC, Bechhoefer J & Rhind N (2015) Replication timing is regulated by the number of MCMs loaded at origins. Genome Res. 25, 1886–1892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Miotto B, Ji Z & Struhl K (2016) Selectivity of ORC binding sites and the relation to replication timing, fragile sites, and deletions in cancers. Proc Natl Acad Sci U S A. 113, E4810–E4819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wang Y, Khan A, Marks AB, Smith OK, Giri S, Lin YC, Creager R, MacAlpine DM, Prasanth KV, Aladjem MI & Prasanth SG (2017) Temporal association of ORCA/LRWD1 to late-firing origins during G1 dictates heterochromatin replication and organization. Nucleic Acids Res. 45, 2490–2502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Sima J, Chakraborty A, Dileep V, Michalski M, Klein KN, Holcomb NP, Turner JL, Paulsen MT, Rivera-Mulia JC, Trevilla-Garcia C, Bartlett DA, Zhao PA, Washburn BK, Nora EP, Kraft K, Mundlos S, Bruneau BG, Ljungman M, Fraser P, Ay F & Gilbert DM (2019) Identifying cis Elements for Spatiotemporal Control of Mammalian DNA Replication. Cell. 176, 816–830 e18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Araki H (2016) Elucidating the DDK-dependent step in replication initiation. EMBO J. 35, 907–908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Parker MW, Botchan MR & Berger JM (2017) Mechanisms and regulation of DNA replication initiation in eukaryotes. Crit Rev Biochem Mol Biol. 52, 107–144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sasi NK, Tiwari K, Soon FF, Bonte D, Wang T, Melcher K, Xu HE & Weinreich M (2014) The potent Cdc7-Dbf4 (DDK) kinase inhibitor XL413 has limited activity in many cancer cell lines and discovery of potential new DDK inhibitor scaffolds. PLoS One. 9, e113300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Zegerman P (2015) Evolutionary conservation of the CDK targets in eukaryotic DNA replication initiation. Chromosoma. 124, 309–321. [DOI] [PubMed] [Google Scholar]
  • 57.Merrick KA & Fisher RP (2010) Putting one step before the other: distinct activation pathways for Cdk1 and Cdk2 bring order to the mammalian cell cycle. Cell Cycle. 9, 706–714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Skaar JR, Pagan JK & Pagano M (2013) Mechanisms and function of substrate recruitment by F-box proteins. Nat Rev Mol Cell Biol. 14, 369–381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Havens CG & Walter JC (2011) Mechanism of CRL4(Cdt2), a PCNA-dependent E3 ubiquitin ligase. Genes Dev. 25, 1568–1582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hayashi A, Giakoumakis NN, Heidebrecht T, Ishii T, Panagopoulos A, Caillat C, Takahara M, Hibbert RG, Suenaga N, Stadnik-Spiewak M, Takahashi T, Shiomi Y, Taraviras S, von Castelmur E, Lygerou Z, Perrakis A & Nishitani H (2018) Direct binding of Cdt2 to PCNA is important for targeting the CRL4(Cdt2) E3 ligase activity to Cdt1. Life Sci Alliance. 1, e201800238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Mazian MA, Suenaga N, Ishii T, Hayashi A, Shiomi Y & Nishitani H (2019) A DNA-binding domain in the C-terminal region of Cdt2 enhances the DNA synthesis-coupled CRL4Cdt2 ubiquitin ligase activity for Cdt1. J Biochem. 165, 505–516. [DOI] [PubMed] [Google Scholar]
  • 62.Starostina NG & Kipreos ET (2012) Multiple degradation pathways regulate versatile CIP/KIP CDK inhibitors. Trends Cell Biol. 22, 33–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Deng T, Yan G, Song X, Xie L, Zhou Y, Li J, Hu X, Li Z, Hu J, Zhang Y, Zhang H, Sun Y, Feng P, Wei D, Hu B, Liu J, Tan W & Ye M (2018) Deubiquitylation and stabilization of p21 by USP11 is critical for cell-cycle progression and DNA damage responses. Proc Natl Acad Sci U S A. 115, 4678–4683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Das CM, Taylor P, Gireud M, Singh A, Lee D, Fuller G, Ji L, Fangusaro J, Rajaram V, Goldman S, Eberhart C & Gopalakrishnan V (2013) The deubiquitylase USP37 links REST to the control of p27 stability and cell proliferation. Oncogene. 32, 1691–1701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Barr AR, Heldt FS, Zhang T, Bakal C & Novak B (2016) A Dynamical Framework for the All-or-None G1/S Transition. Cell Syst. 2, 27–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Cappell SD, Chung M, Jaimovich A, Spencer SL & Meyer T (2016) Irreversible APC(Cdh1) Inactivation Underlies the Point of No Return for Cell-Cycle Entry. Cell. 166, 167–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Grant GD, Kedziora KM, Limas JC, Cook JG & Purvis JE (2018) Accurate delineation of cell cycle phase transitions in living cells with PIP-FUCCI. Cell Cycle. 17, 2496–2516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Spencer SL, Cappell SD, Tsai FC, Overton KW, Wang CL & Meyer T (2013) The proliferation-quiescence decision is controlled by a bifurcation in CDK2 activity at mitotic exit. Cell. 155, 369–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Yuan X, Srividhya J, De Luca T, Lee JH & Pomerening JR (2014) Uncovering the role of APC-Cdh1 in generating the dynamics of S-phase onset. Mol Biol Cell. 25, 441–456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Choudhury R, Bonacci T, Arceci A, Lahiri D, Mills CA, Kernan JL, Branigan TB, DeCaprio JA, Burke DJ & Emanuele MJ (2016) APC/C and SCF(cyclin F) Constitute a Reciprocal Feedback Circuit Controlling S-Phase Entry. Cell Rep. 16, 3359–3372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Grant GD & Cook JG (2017) The Temporal Regulation of S Phase Proteins During G1. Adv Exp Med Biol. 1042, 335–369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kernan J, Bonacci T & Emanuele MJ (2018) Who guards the guardian? Mechanisms that restrain APC/C during the cell cycle. Biochim Biophys Acta Mol Cell Res. 1865, 1924–1933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Cappell SD, Mark KG, Garbett D, Pack LR, Rape M & Meyer T (2018) EMI1 switches from being a substrate to an inhibitor of APC/C(CDH1) to start the cell cycle. Nature. 558, 313–317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Koepp DM, Schaefer LK, Ye X, Keyomarsi K, Chu C, Harper JW & Elledge SJ (2001) Phosphorylation-dependent ubiquitination of cyclin E by the SCFFbw7 ubiquitin ligase. Science. 294, 173–177. [DOI] [PubMed] [Google Scholar]
  • 75.Strohmaier H, Spruck CH, Kaiser P, Won K-A, Sangfelt O & Reed SI (2001) Human F-box protein hCdc4 targets cyclin E for proteolysis and is mutated in a breast cancer cell line. Nature. 413, 316–322. [DOI] [PubMed] [Google Scholar]
  • 76.Katsuno Y, Suzukia A, Sugimurab K, Okumurab K, Zineldeena DH, Shimadaa M, Niidaa H, Mizunoc T, Hanaokac F & Nakanishia M (2009) Cyclin A–Cdk1 regulates the origin firing program in mammalian cells. Proc Natl Acad Sci U S A. 106, 3184–3189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Nakanishi M, Katsuno Y, Niida H, Murakami H & Shimada M (2010) Chk1-cyclin A/Cdk1 axis regulates origin firing programs in mammals. Chromosome Res. 18, 103–113. [DOI] [PubMed] [Google Scholar]
  • 78.Bui DT & Li JJ (2019) DNA Rereplication Is Susceptible to Nucleotide-Level Mutagenesis. Genetics. 212, 445–460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Hanlon SL & Li JJ (2015) Re-replication of a centromere induces chromosomal instability and aneuploidy. PLoS Genet. 11, e1005039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Liontos M, Koutsami M, Sideridou M, Evangelou K, Kletsas D, Levy B, Kotsinas A, Nahum O, Zoumpourlis V, Kouloukoussa M, Lygerou Z, Taraviras S, Kittas C, Bartkova J, Papavassiliou AG, Bartek J, Halazonetis TD & Gorgoulis VG (2007) Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res. 67, 10899–10909. [DOI] [PubMed] [Google Scholar]
  • 81.Munoz S, Bua S, Rodriguez-Acebes S, Megias D, Ortega S, de Martino A & Mendez J (2017) In Vivo DNA Re-replication Elicits Lethal Tissue Dysplasias. Cell Rep. 19, 928–938. [DOI] [PubMed] [Google Scholar]
  • 82.Truong LN & Wu X (2011) Prevention of DNA re-replication in eukaryotic cells. J Mol Cell Biol. 3, 13–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Champeris Tsaniras S, Villiou M, Giannou AD, Nikou S, Petropoulos M, Pateras IS, Tserou P, Karousi F, Lalioti ME, Gorgoulis VG, Patmanidi AL, Stathopoulos GT, Bravou V, Lygerou Z & Taraviras S (2018) Geminin ablation in vivo enhances tumorigenesis through increased genomic instability. J Pathol. 246, 134–140. [DOI] [PubMed] [Google Scholar]
  • 84.Vaziri C, Saxena S, Jeon Y, Lee C, Murata K, Machida Y, Wagle N, Hwang DS & Dutta A (2003) A p53-Dependent Checkpoint Pathway Prevents Rereplication. Mol Cell. 11, 997–1008. [DOI] [PubMed] [Google Scholar]
  • 85.Cook JG, Chasse DA & Nevins JR (2004) The regulated association of Cdt1 with minichromosome maintenance proteins and Cdc6 in mammalian cells. J Biol Chem. 279, 9625–9633. [DOI] [PubMed] [Google Scholar]
  • 86.Pozo PN & Cook JG (2016) Regulation and Function of Cdt1; A Key Factor in Cell Proliferation and Genome Stability. Genes (Basel). 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Yanagi K, Mizuno T, You Z & Hanaoka F (2002) Mouse geminin inhibits not only Cdt1-MCM6 interactions but also a novel intrinsic Cdt1 DNA binding activity. J Biol Chem. 277, 40871–40880. [DOI] [PubMed] [Google Scholar]
  • 88.Ballabeni A, Melixetian M, Zamponi R, Masiero L, Marinoni F & Helin K (2004) Human geminin promotes pre-RC formation and DNA replication by stabilizing CDT1 in mitosis. EMBO J. 23, 3122–3132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Varma D, Chandrasekaran S, Sundin LJ, Reidy KT, Wan X, Chasse DA, Nevis KR, DeLuca JG, Salmon ED & Cook JG (2012) Recruitment of the human Cdt1 replication licensing protein by the loop domain of Hec1 is required for stable kinetochore-microtubule attachment. Nat Cell Biol. 14, 593–603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Machida YJ & Dutta A (2007) The APC/C inhibitor, Emi1, is essential for prevention of rereplication. Genes Dev. 21, 184–194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Bartek J, Lukas C & Lukas J (2004) Checking on DNA damage in S phase. Nat Rev Mol Cell Biol. 5, 792–804. [DOI] [PubMed] [Google Scholar]
  • 92.Li X, Zhao Q, Liao R, Sun P & Wu X (2003) The SCF(Skp2) ubiquitin ligase complex interacts with the human replication licensing factor Cdt1 and regulates Cdt1 degradation. J Biol Chem. 278, 30854–30858. [DOI] [PubMed] [Google Scholar]
  • 93.Liu E, Li X, Yan F, Zhao Q & Wu X (2004) Cyclin-dependent kinases phosphorylate human Cdt1 and induce its degradation. J Biol Chem. 279, 17283–17288. [DOI] [PubMed] [Google Scholar]
  • 94.Yim H, Park JW, Woo SU, Kim ST, Liu L, Lee CH & Lee SK (2013) Phosphorylation of Cdc6 at serine 74, but not at serine 106, drives translocation of Cdc6 to the cytoplasm. J Cell Physiol. 228, 1221–8. [DOI] [PubMed] [Google Scholar]
  • 95.Méndez J, Zou-Yang XH, Kim S-Y, Hidaka M, Tansey WP & Stillman B (2002) Human Origin Recognition Complex Large Subunit Is Degraded by Ubiquitin-Mediated Proteolysis after Initiation of DNA Replication. Molecular Cell. 9, 481–491. [DOI] [PubMed] [Google Scholar]
  • 96.Labib K & De Piccoli G (2011) Surviving chromosome replication: the many roles of the S-phase checkpoint pathway. Philos Trans R Soc Lond B Biol Sci. 366, 3554–3561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Iyer DR & Rhind N (2017) The Intra-S Checkpoint Responses to DNA Damage. Genes (Basel). 8, 1–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Sorensen CS & Syljuasen RG (2012) Safeguarding genome integrity: the checkpoint kinases ATR, CHK1 and WEE1 restrain CDK activity during normal DNA replication. Nucleic Acids Res. 40, 477–486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Daigh LH, Liu C, Chung M, Cimprich KA & Meyer T (2018) Stochastic Endogenous Replication Stress Causes ATR-Triggered Fluctuations in CDK2 Activity that Dynamically Adjust Global DNA Synthesis Rates. Cell Syst. 7, 17–27 e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Eykelenboom JK, Harte EC, Canavan L, Pastor-Peidro A, Calvo-Asensio I, Llorens-Agost M & Lowndes NF (2013) ATR activates the S-M checkpoint during unperturbed growth to ensure sufficient replication prior to mitotic onset. Cell Rep. 5, 1095–1107. [DOI] [PubMed] [Google Scholar]
  • 101.Cayrou C, Coulombe P, Vigneron A, Stanojcic S, Ganier O, Peiffer I, Rivals E, Puy A, Laurent-Chabalier S, Desprat R & Mechali M (2011) Genome-scale analysis of metazoan replication origins reveals their organization in specific but flexible sites defined by conserved features. Genome Res. 21, 1438–1449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Lebofsky R, Heilig R, Sonnleitner M, Weissenbach J & Bensimon A (2006) DNA replication origin interference increases the spacing between initiation events in human cells. Mol Biol Cell. 17, 5337–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Wong PG, Winter SL, Zaika E, Cao TV, Oguz U, Koomen JM, Hamlin JL & Alexandrow MG (2011) Cdc45 limits replicon usage from a low density of preRCs in mammalian cells. PLoS One. 6, 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Woodward AM, Gohler T, Luciani MG, Oehlmann M, Ge X, Gartner A, Jackson DA & Blow JJ (2006) Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress. J Cell Biol. 173, 673–683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Ibarra A, Schwob E & Mendez J (2008) Excess MCM proteins protect human cells from replicative stress by licensing backup origins. Proc Natl Acad Sci U S A. 105, 8956–8961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Marks AB, Smith OK & Aladjem MI (2016) Replication origins: determinants or consequences of nuclear organization? Curr Opin Genet Dev. 37, 67–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Blow JJ, Ge XQ & Jackson DA (2011) How dormant origins promote complete genome replication. Trends Biochem Sci. 36, 405–414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.McIntosh D & Blow JJ (2012) Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb Perspect Biol. 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Ge XQ & Blow JJ (2010) Chk1 inhibits replication factory activation but allows dormant origin firing in existing factories. J Cell Biol. 191, 1285–1297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Shima N, Alcaraz A, Liachko I, Buske TR, Andrews CA, Munroe RJ, Hartford SA, Tye BK & Schimenti JC (2007) A viable allele of Mcm4 causes chromosome instability and mammary adenocarcinomas in mice. Nat Genet. 39, 93–98. [DOI] [PubMed] [Google Scholar]
  • 111.Saldivar JC, Hamperl S, Bocek MJ, Chung M, Bass TE, Cisneros-Soberanis F, Samejima K, Xie L, Paulson JR, Earnshaw WC, Cortez D, Meyer T & Cimprich KA (2018) An intrinsic S/G2 checkpoint enforced by ATR. Science. 361, 806–810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Haland TW, Boye E, Stokke T, Grallert B & Syljuasen RG (2015) Simultaneous measurement of passage through the restriction point and MCM loading in single cells. Nucleic Acids Res. 43, e150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Sakwe AM, Nguyen T, Athanasopoulos V, Shire K & Frappier L (2007) Identification and characterization of a novel component of the human minichromosome maintenance complex. Mol Cell Biol. 27, 3044–3055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Nishiyama A, Frappier L & Mechali M (2011) MCM-BP regulates unloading of the MCM2–7 helicase in late S phase. Genes Dev. 25, 165–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Dewar JM, Low E, Mann M, Raschle M & Walter JC (2017) CRL2(Lrr1) promotes unloading of the vertebrate replisome from chromatin during replication termination. Genes Dev. 31, 275–290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Deng L, Wu RA, Sonneville R, Kochenova OV, Labib K, Pellman D & Walter JC (2019) Mitotic CDK Promotes Replisome Disassembly, Fork Breakage, and Complex DNA Rearrangements. Mol Cell. 73, 915–929 e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Priego Moreno S, Jones RM, Poovathumkadavil D, Scaramuzza S & Gambus A (2019) Mitotic replisome disassembly depends on TRAIP ubiquitin ligase activity. Life Sci Alliance. 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Lengronne A & Pasero P (2014) Closing the MCM cycle at replication termination sites. EMBO Rep. 15, 1226–1227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Bailey R, Priego Moreno S & Gambus A (2015) Termination of DNA replication forks: “Breaking up is hard to do”. Nucleus. 6, 187–196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Ramadan K, Halder S, Wiseman K & Vaz B (2017) Strategic role of the ubiquitin-dependent segregase p97 (VCP or Cdc48) in DNA replication. Chromosoma. 126, 17–32. [DOI] [PubMed] [Google Scholar]

RESOURCES