Pol epsilon is a tetrameric assembly that plays distinct roles during eukaryotic chromosome replication. It catalyses leading strand DNA synthesis; yet this function is dispensable for viability. Its non-catalytic domains instead play an essential role in the assembly of the active replicative helicase and origin activation, while non-essential histone-fold subunits serve a critical function in parental histone redeposition onto newly synthesised DNA. Furthermore, Pol epsilon plays a structural role in linking the RFC–Ctf18 clamp loader to the replisome, supporting processive DNA synthesis, DNA damage response signalling as well as sister chromatid cohesion. In this minireview, we discuss recent biochemical and structural work that begins to explain various aspects of eukaryotic chromosome replication, with a focus on the multiple roles of Pol epsilon in this process.

In all replicating systems a hexameric, ring-shaped helicase is loaded around DNA and uses the energy derived from nucleotide hydrolysis to unwind the double helix, providing the single-stranded DNA template for the replicative polymerases [1]. A direct comparison between the bacterial and eukaryotic systems gives a sense of the higher degree of complexity involved with replicating chromosomes in eukaryotes. Below, we highlight how the leading strand polymerase Pol epsilon is involved in handling this complexity, by participating in several aspects of chromosome replication that are unique to eukaryotes.

In bacteria, the initiator DnaA wraps around the double helix and promotes its melting [2]. DnaA also engages the DnaC loader that in turn promotes the recruitment of the DnaB replicative helicase around single-stranded DNA, in a process that is aided by ATP hydrolysis by DnaB. The ATPase function of DnaC supports DnaB loading, as well as its activation [3–7]. As a result, as soon as DnaB entraps single-stranded DNA, it starts replication fork unwinding [1]. On the contrary, in eukaryotic cells, replicative helicase loading and DNA unwinding are temporally separated [8]. In G1 phase, the initiator, ORC (a DnaA homologue) binds and bends the DNA [9] to load a set of two helicases forming a catalytically inactive head-to-head double hexamer around duplex DNA [10–12], in a process that requires ATP hydrolysis by MCM [13,14]. Activation of DNA unwinding requires the recruitment of a set of firing factors, including Cdc45 and GINS that engage MCM to form the CMG holo-helicase, which both melts and unwinds origin DNA [15–18]. A key role in CMG formation is played by the leading-strand polymerase Pol epsilon, which mediates the recruitment of GINS onto the MCM to promote CMG formation [18–21]. The mechanism of origin activation is only one of many fundamental differences between bacterial and eukaryotic chromosome replication. A second feature that is found in eukaryotes (and most archaeal species), but lacks in bacteria, is the packaging of DNA in nucleosome arrays [22–24]. Nucleosomes must be uncoiled ahead of the replication fork for DNA unwinding to occur, and the evicted parental histones have to be redeposited onto the duplicated DNA, interspersed with newly synthesised histones [25,26]. A key role in the histone redistribution on the two daughter strands is played, again, by Pol epsilon. This function is totally distinct from origin activation and indeed requires different factors within the Pol epsilon protein assembly [27]. A third major difference between bacteria and eukaryotes is the number of dedicated replicative DNA polymerases to extend the leading and lagging strands. While bacteria employ the same DNA Pol III polymerase to synthesise both strands, eukaryotes use two distinct polymerases, with Pol delta discontinuously synthesising Okazaki fragments on the lagging strand and Pol epsilon synthesising the majority of the leading strand template [28–31]. A fourth difference entails the mechanism to achieve processive DNA synthesis. In all domains of life, this involves the deposition of a sliding clamp onto a primer-template junction by the clamp loader. However, while in bacteria one of the clamp loader subunits physically links together the two Pol III polymerases behind the DnaB helicase [32], evidence that the eukaryotic RFC clamp loader is replisome-incorporated still needs to be acquired, although physical contacts with Pol epsilon have been reported for an alternative clamp loader involved in sister chromatid cohesion establishment [33].

In this review article, we describe how recent structural work is shedding light on the disparate functions of Pol epsilon in origin activation, replication fork progression and epigenetic inheritance.

The first step towards the activation of origin-DNA-loaded MCM double hexamers is the phosphorylation of Mcm4 and Mcm6 subunits by the Dbf4-dependent kinase DDK [34–38]. The phosphorylated MCM subunits are recognised by the Sld3 phosphoreader [34], which exists in complex with the Sld7 dimer [39,40], yielding a dimer-of-dimers configuration. Sld3/7 are in turn responsible for the recruitment of the helicase activator, Cdc45 [39,41]. Recruitment of GINS, another helicase-activating factor, depends on a second kinase, CDK, which phosphorylates Sld2, promoting the formation of a so-called pre-loading complex that also contains Dpb11, GINS and Pol epsilon [20]. CDK has a second phosphorylation target amongst firing factors, Sld3. Phospho-Sld2 and phospho-Sld3 bind to a second phosphoreader, Dpb11, which contains a stretch of four BRCT, phospho-peptide binding repeats [42]. As a result of phosphorylation, a so-called pre-initiation super-complex is formed on origin DNA, containing MCM, Sld3/7/Cdc45 and GINS/Pol epsilon/Sld2/Dpb11 [18]. The pre-initiation complex is only loosely associated with the origin, according to in vitro studies. In fact, disassembly can be promoted upon high-salt treatment, which yields the loaded double hexamer scaffold. Tight association of a subset of factors can be achieved upon ATP binding by MCM, which leads to the retention of GINS and Cdc45 on the MCM, in turn causing disengagement of the double hexamer into two separate CMG particles and initial untwisting of duplex DNA. Mcm10 then promotes ejection of the lagging strand template from the MCM central channel, activation of ATP-hydrolysis dependent DNA unwinding and the recruitment of RPA that protects and stabilises the newly established replication fork (Figure 1A) [18].

The role of Pol epsilon in replication origin activation.

Figure 1.
The role of Pol epsilon in replication origin activation.

(A) Schematic representation of the cascade of molecular events leading to the replication fork establishment. (B) The structure of CMG bound to Pol epsilon explains the mechanism whereby Dpb2 and the Pol2 C-terminal domain recruit GINS to the MCM complex.

Figure 1.
The role of Pol epsilon in replication origin activation.

(A) Schematic representation of the cascade of molecular events leading to the replication fork establishment. (B) The structure of CMG bound to Pol epsilon explains the mechanism whereby Dpb2 and the Pol2 C-terminal domain recruit GINS to the MCM complex.

Close modal

Biochemical reconstitution experiments have demonstrated that omission of Pol epsilon prevents CMG formation and origin activation [21], while only a subset of functional domains are required in this process [43,44].

Pol epsilon is a tetrameric assembly composed of a catalytic subunit featuring a tandem repeat of polymerase domains. The first repeat is catalytically active, containing both DNA synthesis and exonuclease functions. The second repeat (C-terminal Pol2 or C-Pol2) has become inactivated during evolution, however, it is the only Pol2 domain essential for viability [45–48]. The second largest Pol epsilon subunit, Dpb2, contains an inactivated calcineurin-like exonuclease domain and is also essential for viability [49]. Finally, the dispensable Dpb3 and Dpb4 subunits contain a histone fold [50], which will be discussed in detail in the next section. C-Pol2 and the Dpb2 are required for CMG formation and DNA replication in vitro, explaining why these two elements are essential for viability [44]. Electron microscopy studies on the CMG and its interaction with Pol epsilon suggest a plausible structural basis for the essential role of C-Pol2/Dpb2 during origin activation. The first report of a CMG structure revealed a topologically closed MCM ring flanked by GINS and Cdc45 that contact the N-terminal domain of MCM and latch across a natural discontinuity in the helicase (Mcm2–5) [17]. Pol epsilon is positioned on the C-terminal, ATPase side of the MCM helicase [51] via its non-catalytic portion [52]. In particular, C-Pol2 contacts Mcm2 and Mcm5 via the catalytically defunct polymerase module and Dpb2 contacts Mcm3 via the inactive exonuclease. As a result, C-Pol2 and Dpb2 close the Mcm2–5 gate on the ATPase side of the helicase motor ring. Dpb2 also contacts the B domain of GINS subunit Psf1, via its N-terminal extension [44], explaining earlier observations that this interaction achieves the integration of Pol epsilon into the replisome and is required for initiation (Figure 1B) [19]. Altogether, the cryo-EM structure of the CMG-Pol epsilon complex invites a model whereby the two essential domains in Pol epsilon serve as a physical link between MCM and GINS, providing an explanation for the structural role of Pol epsilon in CMG formation [44]. It should be noted that the CMG-Pol epsilon assembly was reconstituted in vitro on a model DNA fork, by mixing overexpressed, pre-formed CMG with Pol epsilon in the presence of a nucleotide analogue. Whether the exact same interaction interfaces are recapitulated during origin-dependent CMG formation remains to be established.

Once activated, the CMG helicase must uncoil nucleosomes ahead of the replication fork to achieve the unwinding of parental DNA, and at the same time re-package duplicated DNA filaments into nucleosome arrays [25,26]. Nucleosomes harbour an octameric protein core composed of a pair of four histones (H2A, H2B, H3 and H4), wrapped around a stretch of ∼147 base-pair DNA [53]. Post-translational modification of histones modulates gene expression, encoding an epigenetic programme that is inherited upon cell division. This process maintains cellular fitness and controls harmonious development in multicellular organisms [54].

The mechanism of nucleosome uncoiling by the advancing replisome is not understood. Early work on CMG in the absence of DNA led to the suggestion that the helicase spools DNA through its central pore through a pumpjack-like movement of a set of C-terminal winged-helix domains appended to the MCM module [55]. Later structural work on the DNA-bound complex supports a model whereby the CMG, like other hexameric helicases [56–58], translocates along single-stranded DNA with a hand-over-hand, rotary-cycling mechanism. In this context, neighbouring ATPase sites in MCM ring subunits fire sequentially, causing conformational changes in DNA-interacting pore loops that promote nucleic acid rotation along the inner perimeter of the helicase channel, as well as translation from the N- to C-terminal side of the hexamer [59,60]. The CMG can bypass a roadblock on the lagging but not the leading strand [61,62], implying that the helicase employs a steric exclusion mechanism to split DNA at the fork nexus. A set of pore loops emanating from the N-terminal OB-fold domain of MCM handle the duplex:single-stranded DNA junction, with a conserved phenylalanine in Mcm7 forming pi–pi interactions with the pair base to be melted and a beta hairpin of Mcm3 diverting the lagging strand towards an exit passage formed between Mcm3 and Mcm5 [60,63,64]. Future work will elucidate how duplex-DNA opening is coordinated with nucleosome uncoiling ahead of the fork and which proteins within the replisome assembly facilitate this process. Several elements with histone chaperone function have been identified in core components of the replisome. Amongst these, a negatively charged N-terminal Mcm2 element can wrap around histones H3/H4 protecting a positively charged surface that would be left exposed upon DNA uncoiling [65,66]. An N-terminal domain in the catalytic DNA polymerase subunit of Pol alpha/primase is understood to serve similar functions [67,68]. The histone chaperone FACT (Facilitates Chromosome Transactions) has also been implicated as a replisome component [69–71], although it becomes essential in supporting replication through nucleosomes only when chromatin is densely packed [72,73]. While the role of replisomal histone chaperones in nucleosome uncoiling is unclear, it is established that these factors play a key role in the redeposition of parental histones onto duplicated DNA, which is fundamental for epigenetic inheritance [54]. Selective histone transfer from parental to lagging strand DNA depends on the histone chaperone domain located in the Mcm2 N-terminus, and is likely to occur at the front of the advancing CMG helicase, co-localised with Pol alpha, in turn linked to the CMG via the Ctf4 replisome-organisation hub [51,63,67,74–77]. Histone transfer onto the newly duplicated leading strand DNA depends on Pol epsilon and in particular the Dpb3–Dpb4 subunits [27] (in yeast, or Pole4–Pole3 in humans). In vitro reconstitution work demonstrated that human Pole3–Pole4 function as histone chaperones that can engage H3–H4, promoting tetrasome formation and DNA supercoiling [78] (Figure 2A). The structural homology with histones H2A–H2B [50] invites a tantalising model, whereby Pole3–Pole4 engage a histone H3–H4 tetramer, mimicking its interaction with the H2A–H2B (Figure 2B), and promoting histone redeposition onto the newly synthesised DNA, in a process understood to be supported by FACT [25,26,72]. The full-length structure of yeast Pol epsilon, however, reveals that Dpb3–Dpb4 dimer is clamped between the catalytic lobe and the non-catalytic half of the complex. This architecture confers rigidity to the whole structure [79] (Figure 2C), which is incompatible with a histone-core-like interaction with histones H3–H4 (Figure 2D). A role for Dpb3–Dpb4 (Pole4–Pole3) in the stabilisation of Pol epsilon complex is further supported by work in mice, demonstrating that Pole4 deletion destabilises the whole Pol epsilon complex, leading to embryonic lethality in inbred strains and developmental abnormalities and tumour predisposition in mixed backgrounds [80]. In agreement with the essential role of Pol epsilon in origin activation, Pole4−/− cells showed reduced origin activation and replicative damage. Structural flexibility in the Pol epsilon complex might be required for Dpb3–Dpb4 to engage parental histones. Indeed, several electron microscopy reports indicate that, in particular when engaged with the CMG, Pol epsilon exists primarily in a flexible state, with the catalytic domain of Pol2 free to move with respect to the rest of the complex [44,51,52,81,82]. Such flexibility might render Dpb3–Dpb4 free to engage the H3–H4 parental histones on the path to nucleosome reconstitution on the newly synthesised leading strand DNA by Pol2. Alternatively, Dpb3–Dpb4 engagement with H3–H4 might not involve the histone-like interface per se, but rather unstructured negatively charged tails of Dpb3 and Dpb4. In vitro interaction studies and hydrogen-deuterium exchange mass spectrometry analysis on human proteins support this notion [78]. Indeed, while POLE3-POLE4 and its yeast homologues can separately interact with H3–H4, the negatively charged C-terminal tail of POLE3 is required for the interaction of the POLE3–POLE4 subcomplex with H3–H4 in high salt concentrations [27,77]. Of note, this unstructured domain is predicted to emerge from the histone fold embedded in the core of Pol epsilon, suggesting an important role in histone recycling, likely in cooperation with FACT or other histone chaperones. Further structural analysis is needed to elucidate the mechanism whereby Dpb3–Dpb4 deposit parental H3–H4 onto the newly duplicated leading strand emerging from the N-terminal catalytic domain of Pol2. This work will be also essential to dissect the roles of Pol epsilon instability and defective histone redeposition upon loss of POLE3–POLE4 in mammalian cells [78,80].

The role of Pol epsilon in chromatin replication.

Figure 2.
The role of Pol epsilon in chromatin replication.

(A) Schematic representation of nucleosome uncoiling in front of the replication fork and parental histone redeposition onto duplicated DNA. (B) Modelled interactions between Dpb3–Dpb4 and histones H3–H4 based on the homology with histones H2A–H2B. (C) The structure of full-length Pol epsilon. (D) The rigid structure of full-length Pol epsilon would be able to accommodate an H3–H4 interaction with unstructured Dpb3–Dpb4 tails but not a histone-core-like engagement.

Figure 2.
The role of Pol epsilon in chromatin replication.

(A) Schematic representation of nucleosome uncoiling in front of the replication fork and parental histone redeposition onto duplicated DNA. (B) Modelled interactions between Dpb3–Dpb4 and histones H3–H4 based on the homology with histones H2A–H2B. (C) The structure of full-length Pol epsilon. (D) The rigid structure of full-length Pol epsilon would be able to accommodate an H3–H4 interaction with unstructured Dpb3–Dpb4 tails but not a histone-core-like engagement.

Close modal

Establishment of DNA synthesis upon replication initiation requires the priming of both leading and lagging-strand templates by Pol alpha, with the primase subunit synthesising an RNA oligonucleotide, which is extended by the DNA polymerase subunit Pol1, before the substrate is handed over to processive replicative DNA polymerases [83]. The lagging-strand polymerase Pol delta plays a key role in the establishment of leading strand synthesis [43,84,85]. In fact, mapping of leading strand sites at origins of replication led to the discovery that leading strand synthesis occurring rightward from the origin is established by a lagging strand primer on the left of the origin and vice versa. The primer is extended by Pol delta, before substrate handoff to Pol epsilon. It is this substrate handoff that establishes continuous leading strand extension, differentiating it from discontinuous Okazaki fragment synthesis which will instead occur from the second lagging strand priming event onwards, as the two replisomes migrate in opposed directions [86] (Figure 3A). Whether and to what extent two diverging replisomes remain physically coupled is a matter of debate, with recent cryo-EM work proposing that two CMG helicases might remain associated via Ctf4 during DNA synthesis [76]. As Pol epsilon is bound to the MCM motor of the advancing replisome, structural flexibility is required to support substrate handoff from Pol delta to Pol epsilon. As introduced in the previous section, only the Dpb2 and C-Pol2 domains of the leading strand polymerase have been observed anchored to the CMG helicase, with the N-terminal catalytic domain of Pol2 flexibly tethered to the rest of the complex, in a configuration that would allow rapid substrate engagement (or disengagement, for example in response to DNA damage detection) [44,52]. Whether the CMG-associated Pol epsilon engaged in processive DNA synthesis adopts the rigid configuration recently described for the isolated apo polymerase assembly remains to be determined [79].

Mechanism of leading-strand priming and a role for Ctf18–RFC in establishing processive leading-strand synthesis.

Figure 3.
Mechanism of leading-strand priming and a role for Ctf18–RFC in establishing processive leading-strand synthesis.

(A) Leading strand priming requires substrate hand-off from Pol delta to Pol epsilon. Priming of the leading strand on the leftward moving replisome occurs as the first lagging-strand priming event on the rightward moving replisome. (B) Ctf18–RFC constitutively binds Pol epsilon and loads the PCNA sliding clamp onto leading-strand DNA.

Figure 3.
Mechanism of leading-strand priming and a role for Ctf18–RFC in establishing processive leading-strand synthesis.

(A) Leading strand priming requires substrate hand-off from Pol delta to Pol epsilon. Priming of the leading strand on the leftward moving replisome occurs as the first lagging-strand priming event on the rightward moving replisome. (B) Ctf18–RFC constitutively binds Pol epsilon and loads the PCNA sliding clamp onto leading-strand DNA.

Close modal

Single-molecule TIRF microscopy work on DNA replication reconstituted in vitro provides important knowledge on the compositional dynamics of the yeast replisome during replication fork advancement. For example, Pol delta was found to remain replisome-associated, supporting the synthesis of multiple Okazaki fragments, even when challenged with excess Pol delta in solution [87]. Similar observations have been obtained in single-molecule fluorescence microscopy work on yeast cells [88]. This replisome association depends on an interaction between the Pol32 subunit of the Pol delta and the Pol1 catalytic subunit of Pol alpha [87,89]. The observation that Pol delta remains engaged to the replisome was surprising and different from the commonly accepted notion that Pol delta synthesises Okazaki fragments by migrating away from the fork nexus. Pol epsilon on the other hand, remains stably associated with the CMG during leading strand extension, although it can be exchanged when challenged with excess Pol epsilon in solution [87]. This scenario is distinct from the substrate handoff between different polymerases, which would occur while the same Pol epsilon molecule remains anchored to the CMG [52]. Polymerase exchange has been observed in different replication systems [90–92] and can be explained with the postulation that several weak protein–protein interaction elements connect the polymerase with the rest of the replisome. In the CMG-Pol epsilon complex, these would be represented by the N-terminal domain of Dpb2 that links Pol epsilon to the GINS component of the CMG [19], and by a second interaction interface involving the C-Pol2/Dpb2 domains that contact the MCM ATPase [44]. We note that a change in the DNA engagement and nucleotide-binding state in the CMG can promote a large reconfiguration of the ATPase ring in the MCM [17,55,93], which would reconfigure the largest Pol epsilon interaction interface and could in turn promote polymerase ejection. The observed ability of the eukaryotic replisome to selectively exchange leading-strand polymerase might become critical to restart replication forks after stalling and could promote the recruitment of different polymerases required in the DNA-damage repair process.

Biochemical observations on leading and lagging strand synthesis established that the isolated Pol epsilon is a more processive polymerase compared with Pol delta [94]. Crystallographic analysis of the N-terminal Pol2 catalytic domain explains this observation. Pol2 contains a ‘P-domain’ inserted in the polymerase fold, which is absent from Pol delta and achieves topological encirclement of the DNA substrate [95]. In vitro reconstitution demonstrated that interaction with PCNA (the sliding clamp that tethers the polymerase to the newly synthesised DNA) increases processivity of DNA synthesis both on the leading as well as on the lagging strand [43,94,96,97]. The structure of Pol delta bound to PCNA has been described by cryo-EM of human and yeast ternary complexes that also contain a primer-template junction DNA. Here, the C-terminal domain of the catalytic subunit is anchored to one of three PCNA protomers [98,99]. In this configuration newly synthesised DNA is threaded through PCNA, while other PCNA sites remain free to recruit the FEN1 nuclease for Okazaki fragment maturation [98]. How Pol epsilon contacts PCNA remains to be established [79]. PCNA is loaded onto DNA by the RFC clamp loader (a pentameric ATPase composed of Rfc1–5) [100]. Knowledge on any direct interaction between core replisome components and the clamp loader still needs to be acquired. Given that evidence for stable replisome incorporation of Pol delta has only been recently obtained [87], it is likely that studies on the architecture of the complete lagging-strand replisome might reveal specific interactions with RFC. Likewise, RFC can load PCNA for processive DNA synthesis by Pol epsilon on the leading strand, though direct protein–protein interactions with the leading strand replisome have not been described. However, a heptameric, alternative clamp loader where the Rfc1 subunit is swapped for Ctf18, can form a stable complex with Pol epsilon. Ctf18 exists in a complex with Ctf8 and Dcc1 [101–104], together forming a separate, hook-like module that engages the N-terminal Pol2 catalytic domain of Pol epsilon [33]. Notably, the heptameric Ctf18–Dcc1–Ctf8–RFC complex loads PCNA more efficiently than the pentameric complex missing Dcc1-Ctf8 [104,105]. Ctf18–RFC has been implicated in sister chromatid cohesion [106], checkpoint activation and DNA damage repair. A constitutive interaction between Ctf18–RFC and Pol epsilon suggests that this alternative clamp loader might be a core component of the leading-strand replisome. Although it is not required for viability [107], Ctf18–RFC is therefore well-positioned to protect stalled forks and promote S-phase checkpoint activation (Figure 3B) [33].

Here we reviewed recent advances in biochemistry and structural biology that begin to explain the molecular mechanism of chromosome replication, with a focus on the multiple roles of Pol epsilon in different stages of origin activation and replication fork progression. To reach a complete understanding of the structural mechanism of DNA replication several questions need to be addressed. These include the mechanism of Pol epsilon incorporation into the pre-initiation complex, on the path to replication fork establishment; the mechanism whereby the Dpb3–Dpb4 subunits of Pol epsilon deposit parental histones H3–H4 onto leading strand DNA; a complete view of the leading-strand priming process and substrate handoff between Pol alpha, delta and epsilon, on route to processive and continuous leading strand synthesis; the mechanism whereby the clamp loader hands the clamp-engaged DNA substrate to Pol epsilon to achieve processive leading strand synthesis. Cryo-EM of reconstituted DNA replication reactions will play a major role in elucidating these processes.

  • Biochemical reconstitution and single-particle cryo-EM are shaping our understanding of DNA replication mechanisms.

  • DNA unwinding and nucleosome uncoiling, processive DNA synthesis and the redeposition of parental histones are tightly coupled processes.

  • Future cryo-EM on entire DNA replication reactions will provide a complete understanding of chromosome replication.

The authors declare that there are no competing interests associated with the manuscript.

Work in the E.O.’s laboratory is supported by a Ramon y Cajal fellowship RYC2019-028540-I. Work in R.B.’s laboratory is supported by Barts Charity, CRUK City of London and Royal Society. Work in A.C.’s laboratory is funded jointly by the Wellcome Trust, MRC, and CRUK at the Francis Crick Institute (FC001065) and by the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (grant agreement no. 820102). M.A.C. was supported by a Marie Skłodowska-Curie Action Postdoctoral Fellowship (agreement no. 845939).

Open access for this article was enabled by the participation of The Francis Crick Institute in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.

M.A.C. and A.C. worked on conceptualisation, prepared figures and wrote the manuscript, incorporating sections contributed by E.O. and R.B.

We would like to thank the members of the Costa laboratory for useful discussion.

FACT

Facilitates Chromosome Transactions

1
Bleichert
,
F.
,
Botchan
,
M.R.
and
Berger
,
J.M.
(
2017
)
Mechanisms for initiating cellular DNA replication
.
Science
355
,
eaah6317
2
Duderstadt
,
K.E.
,
Chuang
,
K.
and
Berger
,
J.M.
(
2011
)
DNA stretching by bacterial initiators promotes replication origin opening
.
Nature
478
,
209
213
3
Arias-Palomo
,
E.
,
O'Shea
,
V.L.
,
Hood
,
I.V.
and
Berger
,
J.M.
(
2013
)
The bacterial DnaC helicase loader is a DnaB ring breaker
.
Cell
153
,
438
448
4
Arias-Palomo
,
E.
,
Puri
,
N.
,
O'Shea Murray
,
V.L.
,
Yan
,
Q.
and
Berger
,
J.M.
(
2019
)
Physical basis for the loading of a bacterial replicative helicase onto DNA
.
Mol. Cell
74
,
173
184.e4
5
Puri
,
N.
,
Fernandez
,
A.J.
,
O'Shea Murray
,
V.L.
,
McMillan
,
S.
,
Keck
,
J.L.
and
Berger
,
J.M.
(
2021
)
The molecular coupling between substrate recognition and ATP turnover in a AAA+ hexameric helicase loader
.
eLife
10
,
e64232
6
Mott
,
M.L.
and
Berger
,
J.M.
(
2007
)
DNA replication initiation: mechanisms and regulation in bacteria
.
Nat. Rev. Microbiol.
5
,
343
354
7
Davey
,
M.J.
,
Fang
,
L.
,
McInerney
,
P.
,
Georgescu
,
R.E.
and
O'Donnell
,
M.
(
2002
)
The DnaC helicase loader is a dual ATP/ADP switch protein
.
EMBO J.
21
,
3148
3159
8
Remus
,
D.
and
Diffley
,
J.F.
(
2009
)
Eukaryotic DNA replication control: lock and load, then fire
.
Curr. Opin. Cell Biol.
21
,
771
777
9
Li
,
N.
,
Lam
,
W.H.
,
Zhai
,
Y.
,
Cheng
,
J.
,
Cheng
,
E.
,
Zhao
,
Y.
et al (
2018
)
Structure of the origin recognition complex bound to DNA replication origin
.
Nature
559
,
217
222
10
Remus
,
D.
,
Beuron
,
F.
,
Tolun
,
G.
,
Griffith
,
J.D.
,
Morris
,
E.P.
and
Diffley
,
J.F.
(
2009
)
Concerted loading of Mcm2-7 double hexamers around DNA during DNA replication origin licensing
.
Cell
139
,
719
730
11
Evrin
,
C.
,
Clarke
,
P.
,
Zech
,
J.
,
Lurz
,
R.
,
Sun
,
J.
,
Uhle
,
S.
et al (
2009
)
A double-hexameric MCM2-7 complex is loaded onto origin DNA during licensing of eukaryotic DNA replication
.
Proc. Natl Acad. Sci. U.S.A.
106
,
20240
5
12
Miller
,
T.C.R.
,
Locke
,
J.
,
Greiwe
,
J.F.
,
Diffley
,
J.F.X.
and
Costa
,
A.
(
2019
)
Mechanism of head-to-head MCM double-hexamer formation revealed by cryo-EM
.
Nature
575
,
704
710
13
Coster
,
G.
,
Frigola
,
J.
,
Beuron
,
F.
,
Morris
,
E.P.
and
Diffley
,
J.F.
(
2014
)
Origin licensing requires ATP binding and hydrolysis by the MCM replicative helicase
.
Mol. Cell
55
,
666
677
14
Kang
,
S.
,
Warner
,
M.D.
and
Bell
,
S.P.
(
2014
)
Multiple functions for Mcm2-7 ATPase motifs during replication initiation
.
Mol. Cell
55
,
655
665
15
Moyer
,
S.E.
,
Lewis
,
P.W.
and
Botchan
,
M.R.
(
2006
)
Isolation of the Cdc45/Mcm2-7/GINS (CMG) complex, a candidate for the eukaryotic DNA replication fork helicase
.
Proc. Natl Acad. Sci. U.S.A.
103
,
10236
10241
16
Ilves
,
I.
,
Petojevic
,
T.
,
Pesavento
,
J.J.
and
Botchan
,
M.R.
(
2010
)
Activation of the MCM2-7 helicase by association with Cdc45 and GINS proteins
.
Mol. Cell
37
,
247
258
17
Costa
,
A.
,
Ilves
,
I.
,
Tamberg
,
N.
,
Petojevic
,
T.
,
Nogales
,
E.
,
Botchan
,
M.R.
et al (
2011
)
The structural basis for MCM2-7 helicase activation by GINS and Cdc45
.
Nat. Struct. Mol. Biol.
18
,
471
477
18
Douglas
,
M.E.
,
Ali
,
F.A.
,
Costa
,
A.
and
Diffley
,
J.F.X.
(
2018
)
The mechanism of eukaryotic CMG helicase activation
.
Nature
555
,
265
268
19
Sengupta
,
S.
,
van Deursen
,
F.
,
de Piccoli
,
G.
and
Labib
,
K.
(
2013
)
Dpb2 integrates the leading-strand DNA polymerase into the eukaryotic replisome
.
Curr. Biol.
23
,
543
552
20
Muramatsu
,
S.
,
Hirai
,
K.
,
Tak
,
Y.S.
,
Kamimura
,
Y.
and
Araki
,
H.
(
2010
)
CDK-dependent complex formation between replication proteins Dpb11, Sld2, Pol (epsilon}, and GINS in budding yeast
.
Genes Dev.
24
,
602
612
21
Yeeles
,
J.T.
,
Deegan
,
T.D.
,
Janska
,
A.
,
Early
,
A.
and
Diffley
,
J.F.
(
2015
)
Regulated eukaryotic DNA replication origin firing with purified proteins
.
Nature
519
,
431
435
22
Mattiroli
,
F.
,
Bhattacharyya
,
S.
,
Dyer
,
P.N.
,
White
,
A.E.
,
Sandman
,
K.
,
Burkhart
,
B.W.
et al (
2017
)
Structure of histone-based chromatin in Archaea
.
Science
357
,
609
612
23
Malik
,
H.S.
and
Henikoff
,
S.
(
2003
)
Phylogenomics of the nucleosome
.
Nat. Struct. Mol. Biol.
10
,
882
891
24
Luger
,
K.
,
Mader
,
A.W.
,
Richmond
,
R.K.
,
Sargent
,
D.F.
and
Richmond
,
T.J.
(
1997
)
Crystal structure of the nucleosome core particle at 2.8 A resolution
.
Nature
389
,
251
260
25
Miller
,
T.C.
and
Costa
,
A.
(
2017
)
The architecture and function of the chromatin replication machinery
.
Curr. Opin. Struct. Biol.
47
,
9
16
26
Alabert
,
C.
and
Groth
,
A.
(
2012
)
Chromatin replication and epigenome maintenance
.
Nat. Rev. Mol. Cell Biol.
13
,
153
167
27
Yu
,
C.
,
Gan
,
H.
,
Serra-Cardona
,
A.
,
Zhang
,
L.
,
Gan
,
S.
,
Sharma
,
S.
et al (
2018
)
A mechanism for preventing asymmetric histone segregation onto replicating DNA strands
.
Science
361
,
1386
1389
28
Nick McElhinny
,
S.A.
,
Gordenin
,
D.A.
,
Stith
,
C.M.
,
Burgers
,
P.M.
and
Kunkel
,
T.A.
(
2008
)
Division of labor at the eukaryotic replication fork
.
Mol. Cell
30
,
137
144
29
Pursell
,
Z.F.
,
Isoz
,
I.
,
Lundström
,
E.-B.
,
Johansson
,
E.
and
Kunkel
,
T.A.
(
2007
)
Yeast DNA polymerase ε participates in leading-Strand DNA replication
.
Science
317
,
127
130
30
Yao
,
N.Y.
and
O'Donnell
,
M.
(
2010
)
Snapshot: the replisome
.
Cell
141
,
1088
31
McHenry
,
C.S.
(
2011
)
DNA replicases from a bacterial perspective
.
Annu. Rev. Biochem.
80
,
403
436
32
Xu
,
Z.-Q.
and
Dixon
,
N.E.
(
2018
)
Bacterial replisomes
.
Curr. Opin. Struct. Biol.
53
,
159
168
33
Stokes
,
K.
,
Winczura
,
A.
,
Song
,
B.
,
De Piccoli
,
G.
and
Grabarczyk
,
D.B.
(
2020
)
Ctf18-RFC and DNA Pol ɛ form a stable leading strand polymerase/clamp loader complex required for normal and perturbed DNA replication
.
Nucleic Acids Res.
48
,
8128
8145
34
Deegan
,
T.D.
,
Yeeles
,
J.T.
and
Diffley
,
J.F.
(
2016
)
Phosphopeptide binding by Sld3 links Dbf4-dependent kinase to MCM replicative helicase activation
.
EMBO J.
35
,
961
973
35
Randell
,
J.C.
,
Fan
,
A.
,
Chan
,
C.
,
Francis
,
L.I.
,
Heller
,
R.C.
,
Galani
,
K.
et al (
2010
)
Mec1 is one of multiple kinases that prime the Mcm2-7 helicase for phosphorylation by Cdc7
.
Mol. Cell
40
,
353
363
36
Sheu
,
Y.J.
and
Stillman
,
B.
(
2006
)
Cdc7-Dbf4 phosphorylates MCM proteins via a docking site-mediated mechanism to promote S phase progression
.
Mol. Cell
24
,
101
113
37
Sheu
,
Y.J.
and
Stillman
,
B.
(
2010
)
The Dbf4-Cdc7 kinase promotes S phase by alleviating an inhibitory activity in Mcm4
.
Nature
463
,
113
117
38
Francis
,
L.I.
,
Randell
,
J.C.W.
,
Takara
,
T.J.
,
Uchima
,
L.
and
Bell
,
S.P.
(
2009
)
Incorporation into the prereplicative complex activates the Mcm2–7 helicase for Cdc7–Dbf4 phosphorylation
.
Genes Dev.
23
,
643
654
39
Tanaka
,
T.
,
Umemori
,
T.
,
Endo
,
S.
,
Muramatsu
,
S.
,
Kanemaki
,
M.
,
Kamimura
,
Y.
et al (
2011
)
Sld7, an Sld3-associated protein required for efficient chromosomal DNA replication in budding yeast
.
EMBO J.
30
,
2019
2030
40
Itou
,
H.
,
Shirakihara
,
Y.
and
Araki
,
H.
(
2015
)
The quaternary structure of the eukaryotic DNA replication proteins Sld7 and Sld3
.
Acta Crystallogr. D Biol. Crystallogr.
71
,
1649
1656
41
Kamimura
,
Y.
,
Tak
,
Y.S.
,
Sugino
,
A.
and
Araki
,
H.
(
2001
)
Sld3, which interacts with Cdc45 (Sld4), functions for chromosomal DNA replication in Saccharomyces cerevisiae
.
EMBO J.
20
,
2097
2107
42
Zegerman
,
P.
and
Diffley
,
J.F.
(
2007
)
Phosphorylation of Sld2 and Sld3 by cyclin-dependent kinases promotes DNA replication in budding yeast
.
Nature
445
,
281
285
43
Yeeles
,
J.T.
,
Janska
,
A.
,
Early
,
A.
and
Diffley
,
J.F.
(
2017
)
How the eukaryotic replisome achieves rapid and efficient DNA replication
.
Mol. Cell
65
,
105
116
44
Goswami
,
P.
,
Abid Ali
,
F.
,
Douglas
,
M.E.
,
Locke
,
J.
,
Purkiss
,
A.
,
Janska
,
A.
et al (
2018
)
Structure of DNA-CMG-Pol epsilon elucidates the roles of the non-catalytic polymerase modules in the eukaryotic replisome
.
Nat. Commun.
9
,
5061
45
Tahirov
,
T.H.
,
Makarova
,
K.S.
,
Rogozin
,
I.B.
,
Pavlov
,
Y.I.
and
Koonin
,
E.V.
(
2009
)
Evolution of DNA polymerases: an inactivated polymerase-exonuclease module in Pol epsilon and a chimeric origin of eukaryotic polymerases from two classes of archaeal ancestors
.
Biol. Direct
4
,
11
46
Johansson
,
E.
and
Macneill
,
S.A.
(
2010
)
The eukaryotic replicative DNA polymerases take shape
.
Trends Biochem. Sci.
35
,
339
347
47
Dua
,
R.
,
Levy
,
D.L.
and
Campbell
,
J.L.
(
1998
)
Role of the putative zinc finger domain of Saccharomyces cerevisiae DNA polymerase epsilon in DNA replication and the S/M checkpoint pathway
.
J. Biol. Chem.
273
,
30046
30055
48
Feng
,
W.
and
D'Urso
,
G.
(
2001
)
Schizosaccharomyces pombe cells lacking the amino-terminal catalytic domains of DNA polymerase epsilon are viable but require the DNA damage checkpoint control
.
Mol. Cell. Biol.
21
,
4495
4504
49
Baranovskiy
,
A.G.
,
Gu
,
J.
,
Babayeva
,
N.D.
,
Kurinov
,
I.
,
Pavlov
,
Y.I.
and
Tahirov
,
T.H.
(
2017
)
Crystal structure of the human Pol B-subunit in complex with the C-terminal domain of the catalytic subunit
.
J. Biol. Chem.
292
,
15717
15730
50
He
,
H.
,
Li
,
Y.
,
Dong
,
Q.
,
Chang
,
A.Y.
,
Gao
,
F.
,
Chi
,
Z.
et al (
2017
)
Coordinated regulation of heterochromatin inheritance by Dpb3-Dpb4 complex
.
Proc. Natl Acad. Sci. U.S.A.
114
,
12524
12529
51
Sun
,
J.
,
Shi
,
Y.
,
Georgescu
,
R.E.
,
Yuan
,
Z.
,
Chait
,
B.T.
,
Li
,
H.
et al (
2015
)
The architecture of a eukaryotic replisome
.
Nat. Struct. Mol. Biol.
22
,
976
982
52
Zhou
,
J.C.
,
Janska
,
A.
,
Goswami
,
P.
,
Renault
,
L.
,
Abid Ali
,
F.
,
Kotecha
,
A.
et al (
2017
)
CMG-Pol epsilon dynamics suggests a mechanism for the establishment of leading-strand synthesis in the eukaryotic replisome
.
Proc. Natl Acad. Sci. U.S.A.
114
,
4141
4146
53
Zhou
,
K.
,
Gaullier
,
G.
and
Luger
,
K.
(
2019
)
Nucleosome structure and dynamics are coming of age
.
Nat. Struct. Mol. Biol.
26
,
3
13
54
Stewart-Morgan
,
K.R.
,
Petryk
,
N.
and
Groth
,
A.
(
2020
)
Chromatin replication and epigenetic cell memory
.
Nat. Cell Biol.
22
,
361
371
55
Yuan
,
Z.
,
Bai
,
L.
,
Sun
,
J.
,
Georgescu
,
R.
,
Liu
,
J.
,
O'Donnell
,
M.E.
et al (
2016
)
Structure of the eukaryotic replicative CMG helicase suggests a pumpjack motion for translocation
.
Nat. Struct. Mol. Biol.
23
,
217
224
56
Thomsen
,
N.D.
and
Berger
,
J.M.
(
2009
)
Running in reverse: the structural basis for translocation polarity in hexameric helicases
.
Cell
139
,
523
534
57
Enemark
,
E.J.
and
Joshua-Tor
,
L.
(
2006
)
Mechanism of DNA translocation in a replicative hexameric helicase
.
Nature
442
,
270
275
58
Gao
,
Y.
,
Cui
,
Y.
,
Fox
,
T.
,
Lin
,
S.
,
Wang
,
H.
,
de Val
,
N.
et al (
2019
)
Structures and operating principles of the replisome
.
Science
363
,
eaav7003
59
Meagher
,
M.
,
Epling
,
L.B.
and
Enemark
,
E.J.
(
2019
)
DNA translocation mechanism of the MCM complex and implications for replication initiation
.
Nat. Commun.
10
,
3117
60
Eickhoff
,
P.
,
Kose
,
H.B.
,
Martino
,
F.
,
Petojevic
,
T.
,
Abid Ali
,
F.
,
Locke
,
J.
et al (
2019
)
Molecular basis for ATP-hydrolysis-driven DNA translocation by the CMG helicase of the eukaryotic replisome
.
Cell Rep.
28
,
2673
2688 e8
61
Kose
,
H.B.
,
Larsen
,
N.B.
,
Duxin
,
J.P.
and
Yardimci
,
H.
(
2019
)
Dynamics of the eukaryotic replicative helicase at lagging-strand protein barriers support the steric exclusion model
.
Cell Rep.
26
,
2113
2125.e6
62
Langston
,
L.D.
,
Mayle
,
R.
,
Schauer
,
G.D.
,
Yurieva
,
O.
,
Zhang
,
D.
,
Yao
,
N.Y.
et al (
2017
)
Mcm10 promotes rapid isomerization of CMG-DNA for replisome bypass of lagging strand DNA blocks
.
eLife
6
,
e29118
63
Baretić
,
D.
,
Jenkyn-Bedford
,
M.
,
Aria
,
V.
,
Cannone
,
G.
,
Skehel
,
M.
and
Yeeles
,
J.T.P.
(
2020
)
Cryo-EM structure of the fork protection complex bound to CMG at a replication fork
.
Mol. Cell
78
,
926
940.e13
64
Yuan
,
Z.
,
Georgescu
,
R.
,
Bai
,
L.
,
Zhang
,
D.
,
Li
,
H.
and
O'Donnell
,
M.E.
(
2020
)
DNA unwinding mechanism of a eukaryotic replicative CMG helicase
.
Nat. Commun.
11
,
688
65
Richet
,
N.
,
Liu
,
D.
,
Legrand
,
P.
,
Velours
,
C.
,
Corpet
,
A.
,
Gaubert
,
A.
et al (
2015
)
Structural insight into how the human helicase subunit MCM2 may act as a histone chaperone together with ASF1 at the replication fork
.
Nucleic Acids Res.
43
,
1905
1917
66
Huang
,
H.
,
Stromme
,
C.B.
,
Saredi
,
G.
,
Hodl
,
M.
,
Strandsby
,
A.
,
Gonzalez-Aguilera
,
C.
et al (
2015
)
A unique binding mode enables MCM2 to chaperone histones H3-H4 at replication forks
.
Nat. Struct. Mol. Biol.
22
,
618
626
67
Gan
,
H.
,
Serra-Cardona
,
A.
,
Hua
,
X.
,
Zhou
,
H.
,
Labib
,
K.
,
Yu
,
C.
et al (
2018
)
The Mcm2-Ctf4-Polalpha axis facilitates parental histone H3-H4 transfer to lagging strands
.
Mol. Cell
72
,
140
151.e3
68
Evrin
,
C.
,
Maman
,
J.D.
,
Diamante
,
A.
,
Pellegrini
,
L.
and
Labib
,
K.
(
2018
)
Histone H2A-H2B binding by Pol alpha in the eukaryotic replisome contributes to the maintenance of repressive chromatin
.
EMBO J.
37
,
e99021
69
Wittmeyer
,
J.
and
Formosa
,
T.
(
1997
)
The saccharomyces cerevisiae DNA polymerase alpha catalytic subunit interacts with Cdc68/Spt16 and with Pob3, a protein similar to an HMG1-like protein
.
Mol. Cell. Biol.
17
,
4178
4190
70
VanDemark
,
A.P.
,
Blanksma
,
M.
,
Ferris
,
E.
,
Heroux
,
A.
,
Hill
,
C.P.
and
Formosa
,
T.
(
2006
)
The structure of the yFACT Pob3-M domain, its interaction with the DNA replication factor RPA, and a potential role in nucleosome deposition
.
Mol. Cell
22
,
363
374
71
Tan
,
B.C.
,
Chien
,
C.T.
,
Hirose
,
S.
and
Lee
,
S.C.
(
2006
)
Functional cooperation between FACT and MCM helicase facilitates initiation of chromatin DNA replication
.
EMBO J.
25
,
3975
3985
72
Kurat
,
C.F.
,
Yeeles
,
J.T.
,
Patel
,
H.
,
Early
,
A.
and
Diffley
,
J.F.
(
2017
)
Chromatin controls DNA replication origin selection, lagging-strand synthesis, and replication fork rates
.
Mol. Cell
65
,
117
130
73
Devbhandari
,
S.
,
Jiang
,
J.
,
Kumar
,
C.
,
Whitehouse
,
I.
and
Remus
,
D.
(
2017
)
Chromatin constrains the initiation and elongation of DNA replication
.
Mol. Cell
65
,
131
141
74
Simon
,
A.C.
,
Zhou
,
J.C.
,
Perera
,
R.L.
,
van Deursen
,
F.
,
Evrin
,
C.
,
Ivanova
,
M.E.
et al (
2014
)
A Ctf4 trimer couples the CMG helicase to DNA polymerase alpha in the eukaryotic replisome
.
Nature
510
,
293
297
75
Petryk
,
N.
,
Dalby
,
M.
,
Wenger
,
A.
,
Stromme
,
C.B.
,
Strandsby
,
A.
,
Andersson
,
R.
et al (
2018
)
MCM2 promotes symmetric inheritance of modified histones during DNA replication
.
Science
361
,
1389
1392
76
Yuan
,
Z.
,
Georgescu
,
R.
,
Santos
,
R.L.A.
,
Zhang
,
D.
,
Bai
,
L.
,
Yao
,
N.Y.
et al (
2019
)
Ctf4 organizes sister replisomes and Pol alpha into a replication factory
.
eLife
8
,
e47405
77
Rzechorzek
,
N.J.
,
Hardwick
,
S.W.
,
Jatikusumo
,
V.A.
,
Chirgadze
,
D.Y.
and
Pellegrini
,
L.
(
2020
)
CryoEM structures of human CMG–ATPγS–DNA and CMG–AND-1 complexes
.
Nucleic Acids Res.
48
,
6980
6995
78
Bellelli
,
R.
,
Belan
,
O.
,
Pye
,
V.E.
,
Clement
,
C.
,
Maslen
,
S.L.
,
Skehel
,
J.M.
et al (
2018
)
POLE3-POLE4 is a histone H3-H4 chaperone that maintains chromatin integrity during DNA replication
.
Mol. Cell
72
,
112
126.e5
79
Yuan
,
Z.
,
Georgescu
,
R.
,
Schauer
,
G.D.
,
O'Donnell
,
M.E.
and
Li
,
H.
(
2020
)
Structure of the polymerase ε holoenzyme and atomic model of the leading strand replisome
.
Nat. Commun.
11
,
3156
80
Bellelli
,
R.
,
Borel
,
V.
,
Logan
,
C.
,
Svendsen
,
J.
,
Cox
,
D.E.
,
Nye
,
E.
et al (
2018
)
Polε instability drives replication stress, abnormal development, and tumorigenesis
.
Mol. Cell
70
,
707
721.e7
81
Asturias
,
F.J.
,
Cheung
,
I.K.
,
Sabouri
,
N.
,
Chilkova
,
O.
,
Wepplo
,
D.
and
Johansson
,
E.
(
2006
)
Structure of saccharomyces cerevisiae DNA polymerase epsilon by cryo-electron microscopy
.
Nat. Struct. Mol. Biol.
13
,
35
43
82
Jones
,
M.L.
,
Baris
,
Y.
,
Taylor
,
M.R.G.
and
Yeeles
,
J.T.P.
(
2021
)
Structure of a human replisome shows the organisation and interactions of a DNA replication machine
.
EMBO J.
40
,
e108819
83
Pellegrini
,
L.
(
2012
)
The Pol α-primase complex
.
Subcellular Biochem.
62
,
157
169
84
Zhou
,
Z.-X.
,
Lujan
,
S.A.
,
Burkholder
,
A.B.
,
Garbacz
,
M.A.
and
Kunkel
,
T.A.
(
2019
)
Roles for DNA polymerase δ in initiating and terminating leading strand DNA replication
.
Nat. Commun.
10
,
3992
85
Garbacz
,
M.A.
,
Lujan
,
S.A.
,
Burkholder
,
A.B.
,
Cox
,
P.B.
,
Wu
,
Q.
,
Zhou
,
Z.-X.
et al (
2018
)
Evidence that DNA polymerase δ contributes to initiating leading strand DNA replication in saccharomyces cerevisiae
.
Nat. Commun.
9
,
858
86
Aria
,
V.
and
Yeeles
,
J.T.
(
2018
)
Mechanism of bidirectional leading-strand synthesis establishment at eukaryotic DNA replication origins
.
Mol. Cell
73
,
199
211.e10
87
Lewis
,
J.S.
,
Spenkelink
,
L.M.
,
Schauer
,
G.D.
,
Yurieva
,
O.
,
Mueller
,
S.H.
,
Natarajan
,
V.
et al (
2020
)
Tunability of DNA polymerase stability during eukaryotic DNA replication
.
Mol. Cell
77
,
17
25.e5
88
Kapadia
,
N.
,
El-Hajj
,
Z.W.
,
Zheng
,
H.
,
Beattie
,
T.R.
,
Yu
,
A.
and
Reyes-Lamothe
,
R.
(
2020
)
Processive activity of replicative DNA polymerases in the replisome of live eukaryotic cells
.
Mol. Cell
80
,
114
126.e8
89
Johansson
,
E.
,
Garg
,
P.
and
Burgers
,
P.M.
(
2004
)
The Pol32 subunit of DNA polymerase delta contains separable domains for processive replication and proliferating cell nuclear antigen (PCNA) binding
.
J. Biol. Chem.
279
,
1907
1915
90
Aberg
,
C.
,
Duderstadt
,
K.E.
and
van Oijen
,
A.M.
(
2016
)
Stability versus exchange: a paradox in DNA replication
.
Nucleic Acids Res.
44
,
4846
4854
91
Lewis
,
J.S.
,
Spenkelink
,
L.M.
,
Jergic
,
S.
,
Wood
,
E.A.
,
Monachino
,
E.
,
Horan
,
N.P.
et al (
2017
)
Single-molecule visualization of fast polymerase turnover in the bacterial replisome
.
eLife
6
,
e23932
92
Beattie
,
T.R.
,
Kapadia
,
N.
,
Nicolas
,
E.
,
Uphoff
,
S.
,
Wollman
,
A.J.
,
Leake
,
M.C.
et al (
2017
)
Frequent exchange of the DNA polymerase during bacterial chromosome replication
.
eLife
6
,
e21763
93
Abid Ali
,
F.
,
Renault
,
L.
,
Gannon
,
J.
,
Gahlon
,
H.L.
,
Kotecha
,
A.
,
Zhou
,
J.C.
et al (
2016
)
Cryo-EM structures of the eukaryotic replicative helicase bound to a translocation substrate
.
Nat. commun.
7
,
10708
94
Chilkova
,
O.
,
Stenlund
,
P.
,
Isoz
,
I.
,
Stith
,
C.M.
,
Grabowski
,
P.
,
Lundström
,
E.-B.
et al (
2007
)
The eukaryotic leading and lagging strand DNA polymerases are loaded onto primer-ends via separate mechanisms but have comparable processivity in the presence of PCNA
.
Nucleic Acids Res.
35
,
6588
6597
95
Hogg
,
M.
,
Osterman
,
P.
,
Bylund
,
G.O.
,
Ganai
,
R.A.
,
Lundstrom
,
E.B.
,
Sauer-Eriksson
,
A.E.
et al (
2014
)
Structural basis for processive DNA synthesis by yeast DNA polymerase varepsilon
.
Nat. Struct. Mol. Biol.
21
,
49
55
96
Georgescu
,
R.E.
,
Schauer
,
G.D.
,
Yao
,
N.Y.
,
Langston
,
L.D.
,
Yurieva
,
O.
,
Zhang
,
D.
et al (
2015
)
Reconstitution of a eukaryotic replisome reveals suppression mechanisms that define leading/lagging strand operation
.
eLife
4
,
e04988
97
Georgescu
,
R.E.
,
Langston
,
L.
,
Yao
,
N.Y.
,
Yurieva
,
O.
,
Zhang
,
D.
,
Finkelstein
,
J.
et al (
2014
)
Mechanism of asymmetric polymerase assembly at the eukaryotic replication fork
.
Nat. Struct. Mol. Biol.
21
,
664
670
98
Lancey
,
C.
,
Tehseen
,
M.
,
Raducanu
,
V.-S.
,
Rashid
,
F.
,
Merino
,
N.
,
Ragan
,
T.J.
et al (
2020
)
Structure of the processive human Pol δ holoenzyme
.
Nat. Commun.
11
,
1109
99
Zheng
,
F.
,
Georgescu
,
R.E.
,
Li
,
H.
and
O'Donnell
,
M.E.
(
2020
)
Structure of eukaryotic DNA polymerase δ bound to the PCNA clamp while encircling DNA
.
Proc. Natl Acad. Sci. U.S.A.
117
,
30344
30353
100
Odonnell
,
M.
and
Kuriyan
,
J.
(
2006
)
Clamp loaders and replication initiation
.
Curr. Opin. Struct. Biol.
16
,
35
41
101
Murakami
,
T.
,
Takano
,
R.
,
Takeo
,
S.
,
Taniguchi
,
R.
,
Ogawa
,
K.
,
Ohashi
,
E.
et al (
2010
)
Stable interaction between the human proliferating cell nuclear antigen loader complex Ctf18-replication factor C (RFC) and DNA polymerase ɛ is mediated by the cohesion-specific subunits, Ctf18, Dcc1, and Ctf8*
.
J. Biol. Chem.
285
,
34608
34615
102
Okimoto
,
H.
,
Tanaka
,
S.
,
Araki
,
H.
,
Ohashi
,
E.
and
Tsurimoto
,
T.
(
2016
)
Conserved interaction of Ctf18-RFC with DNA polymerase ε is critical for maintenance of genome stability in saccharomyces cerevisiae
.
Genes Cells
21
,
482
491
103
García-Rodríguez
,
L.J.
,
De Piccoli
,
G.
,
Marchesi
,
V.
,
Jones
,
R.C.
,
Edmondson
,
R.D.
and
Labib
,
K.
(
2015
)
A conserved Polɛ binding module in Ctf18-RFC is required for S-phase checkpoint activation downstream of Mec1
.
Nucleic Acids Res.
43
,
8830
8838
104
Grabarczyk
,
D.B.
,
Silkenat
,
S.
and
Kisker
,
C.
(
2018
)
Structural basis for the recruitment of Ctf18-RFC to the replisome
.
Structure
26
,
137
144.e3
105
Fujisawa
,
R.
,
Ohashi
,
E.
,
Hirota
,
K.
and
Tsurimoto
,
T.
(
2017
)
Human CTF18-RFC clamp-loader complexed with non-synthesising DNA polymerase ε efficiently loads the PCNA sliding clamp
.
Nucleic Acids Res.
45
,
4550
4563
106
Mayer
,
M.L.
,
Gygi
,
S.P.
,
Aebersold
,
R.
and
Hieter
,
P.
(
2001
)
Identification of RFC(Ctf18p, Ctf8p, Dcc1p): an alternative RFC complex required for sister chromatid cohesion in S. cerevisiae
.
Mol. Cell
7
,
959
970
107
Kim
,
J.
,
Robertson
,
K.
,
Mylonas
,
K.J.L.
,
Gray
,
F.C.
,
Charapitsa
,
I.
and
MacNeill
,
S.A.
(
2005
)
Contrasting effects of Elg1–RFC and Ctf18–RFC inactivation in the absence of fully functional RFC in fission yeast
.
Nucleic Acids Res.
33
,
4078
4089
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).