Abstract
Eukaryotic replication proteins are highly conserved, and thus study of Saccharomyces cerevisiae replication can inform about this central process in higher eukaryotes including humans. The S. cerevisiae replisome is a large and dynamic assembly comprised of ~50 proteins. The core of the replisome is composed of 31 different proteins including the 11-subunit CMG helicase; RFC clamp loader pentamer; PCNA clamp; the heteroligomeric DNA polymerases ε, δ, and α-primase; and the RPA heterotrimeric single strand binding protein. Many additional protein factors either travel with or transiently associate with these replisome proteins at particular times during replication. In this chapter, we summarize several recent structural studies on the S. cerevisiae replisome and its subassemblies using single particle electron microscopy and X-ray crystallography. These recent structural studies have outlined the overall architecture of a core replisome subassembly and shed new light on the mechanism of eukaryotic replication.
Keywords: Eukaryotic DNA replication, Replisome, Replicative helicase, Cryo-EM, Mcm, CMG, DNA polymerase, Structural biology
10.1 Introduction
The DNA genome is duplicated in semiconservative fashion wherein each strand serves as the template for synthesis of its complementary strand (Meselson and Stahl 1958; Watson and Crick 1953a, b). This process may seem simple, but nothing could be further from the truth; it requires >50 different proteins in eukaryotes, and many of their individual functions remain unknown, much less how they work together to accomplish high-fidelity genome duplication (Costa et al. 2013; Li and Araki 2013; MacNeill 2012). Unlike protein synthesis and mRNA synthesis that are well conserved across the three domains of life, DNA synthesis is not well conserved. Many key replication proteins, such as the replicative helicase, DNA polymerases, primase, and single strand binding protein, are evolutionarily unrelated between bacteria and archaea/eukarya (Forterre et al. 2004; Leipe et al. 1999). Thus one cannot rely on the wealth of bacterial studies to understand eukaryotic replication, and detailed studies of eukaryotic replication mechanisms are warranted. Fortunately, the replication machinery within eukaryotes is highly conserved, and thus the relatively simple Saccharomyces cerevisiae can be used as a reliable model for replication in higher eukaryotes (Leman and Noguchi 2013).
The “core replisome” encompasses proteins that are necessary to propagate the replication fork (Zhang and O’Donnell 2016). The eukaryotic replisome core is composed of CMG helicase, the leading strand DNA polymerase (Pol) ε, lagging strand Pol δ, Pol α-primase, along with the RFC clamp loader, PCNA clamp, and RPA single strand (ss) DNA-binding protein. Many other components, such as Ctf4, Mcm10, Topo I, the checkpoint response factors Mrc1, Tof1, Csm3, and the FACT nucleosome mobility factor, either travel with the replisome or hop on and off at different points during replication (Bell and Labib 2016).
The eukaryotic helicase is an 11-protein machine composed of Cdc45, the Mcm2-7 heterohexamer, and the four-protein GINS complex (Psf1-3, and Sld5) referred to as CMG (Bochman et al. 2008; Ilves et al. 2010; Moyer et al. 2006). Eukaryotic CMG helicase is assembled through a series of replication initiation events (reviewed in (Costa et al. 2013; O’Donnell et al. 2013; Tanaka and Araki 2013; Tognetti et al. 2015)). Briefly, the origin recognition complex (ORC) along with Cdc6 loads onto replication origins in the G1 phase of the cell cycle (Bell and Stillman 1992) and then recruits the helicase core Mcm2-7 heterohexamer with the help of Cdt1 (Cocker et al. 1996; Liang et al. 1995; Mimura et al. 2004; Nishitani et al. 2000; Santocanale and Diffley 1996; Sun et al. 2013; Tanaka and Diffley 2002). Subsequently, another Mcm2-7 is recruited onto double-stranded (ds) DNA forming a double hexamer of Mcm2-7 that is inactive (Duzdevich et al. 2015; Evrin et al. 2009; Remus et al. 2009; Sun et al. 2014; Ticau et al. 2015). Finally, at the G1-to-S transition, the double hexamers are converted to two active helicases comprised of Cdc45-Mcm2-7-GINS (CMG) that each encircle opposite strands of single-stranded (ss) DNA to carry out bidirectional separation of the DNA duplex (Bell and Botchan 2013; Duzdevich et al. 2015).
Eukaryotic replication requires at least three DNA polymerases to propagate fork movement (Burgers 2009). These include Pol α-primase, Pol ε, and Pol δ. Pols α, ε, and δ are all members of the B family of DNA polymerases (Steitz 1999). Pol α initiates DNA synthesis on both the leading and lagging strands by synthesizing a RNA/DNA hybrid primer (Conaway and Lehman 1982). Numerous genetic, cell biology, and biochemical studies concur that Pol ε and Pol δ extend primers on the leading and lagging strands, respectively (Burgers et al. 2016; Clausen et al. 2015; Kunkel and Burgers 2008; Miyabe et al. 2011; Nick McElhinny et al. 2008; Pursell et al. 2007), although one report suggests that Pol δ functions on both strands leaving some uncertainty for future studies to resolve (Johnson et al. 2015).
The antiparallel structure of dsDNA, combined with unidirectional synthesis by DNA polymerases, results in continuous synthesis on the leading strand that travels in the direction of DNA unwinding and discontinuous synthesis on the lagging strand that is extended as multiple Okazaki fragments in the direction opposite DNA unwinding. A preformed primed site is required for DNA polymerase function, and both strands are primed by Pol α-primase. However, Pol α-primase is needed repeatedly on the lagging strand, once for each 100–200 bp Okazaki fragment. Pol α-primase is a four-subunit enzyme; the largest subunit is a DNA polymerase that lacks a proofreading 3′–5′ exonuclease, the second largest is the B subunit (unknown function), and the two small subunits, PriL and PriS, contain the priming activity (Banks et al. 1979; Conaway and Lehman 1982; Kaguni et al. 1983a, b).
There have been numerous important advances on the structure of components of the eukaryotic replisome in the last decade. In this chapter, we focus mainly on structures that have been solved during the past 3 years, specifically of proteins directly involved in moving replication forks. We first review the cryo-EM structure of the CMG helicase and compare the structures of CMG components GINS and Cdc45 of yeast and human. We then review the relative positions of Pol ε, Ctf4, and Pol α in the context of the CMG helicase and the possible replisome architecture at the replication fork. We end our chapter by providing a brief perspective on key missing information and what to expect in the coming years.
10.2 The CMG Structure
The first structure of eukaryotic CMG was determined in the Drosophila melanogaster system by negative stain EM at low resolution (Costa et al. 2011). The structure revealed a Mcm2-7 ring braced on the side by Cdc45 and GINS that interacted mainly with Mcm2, Mcm3, and Mcm5, apparently forming a secondary pore. However, this secondary pore is essentially filled in by side chains in the higher resolution cryo-EM map of the S. cerevisiae CMG at a resolution of 3.7–4.8 Å (Fig. 10.1a–c) (Yuan et al. 2016). The secondary pore also disappears in a medium resolution cryo-EM map of Drosophila CMG at 7–10 Å resolution (Abid Ali et al. 2016). The Mcm2-7 core in the CMG forms a two-tiered ring structure: an N-terminal domain (NTD) tier composed of a helical subdomain, a Zn-binding motif, and an oligonucleotide/oligosaccharide-binding (OB) motif, while the C-terminal domain (CTD) tier contains the AAA+ motors. This domain architecture is similar to that of other hexameric helicases including the archaeal MCM hexamer and the inactive yeast Mcm2-7 in the double hexamer (Brewster et al. 2008; Enemark and Joshua-Tor 2006; Li et al. 2015; Miller et al. 2014; Singleton et al. 2000; Slaymaker and Chen 2012). Cdc45 and GINS mainly bind the NTD-tier ring of Mcm2-7, forming a rigid unit of Cdc45-GINS-Mcm2-7 NTD. More precisely, Cdc45 contacts only the NTDs of Mcm2 and Mcm5, and GINS contacts only the NTDs of Mcm3 and Mcm5 (GINS and Cdc45 also contact one another). It is the tight interaction between the NTD of Mcm5 and Cdc45-GINS that occludes the secondary pore.
In the medium resolution cryo-EM structures of the Drosophila CMG, density corresponding to six nucleotides of ssDNA was observed either inside the CTD-tier motor ring or inside the NTD-tier ring (Abid Ali et al. 2016). This observation is consistent with CMG acting as an ssDNA translocase. Interestingly, in an archaeal MCM hexamer, ssDNA was found to bind the NTD-tier ring as a circle perpendicular to the MCM channel axis and was proposed to be an intermediate in origin initiation (Froelich et al. 2014). How a CMG engages DNA at a replication fork remains an important issue in the field.
10.3 The Mcm2-7 Hexamer Undergoes Large Conformational Changes from the Inactive Double Hexamer to the Active CMG Helicase
Recent studies have provided two yeast Mcm2-7 hexamer structures: one determined in the context of the inactive Mcm2-7 double hexamer and the other in the context of active CMG as described above (Li et al. 2015; Yuan et al. 2016). This allows a comparison of the structures to gain insight into the mechanism by which the helicase is activated. To transition from the inactive Mcm2-7 double hexamer to the single Mcm2-7 in CMG requires large domain movements of up to 20 Å and rotations of up to 30° (Yuan et al. 2016). Movements in the CTD-tier motor domains are more extensive than the NTD-tier ring. As an example, we compare Mcm2-Mcm5 in the two structures (Fig. 10.2a–c). Notably, the interface between the CTDs of Mcm2 and Mcm5 is looser in the CMG structure than in the double hexamer. Activation of CMG requires it to encircle the leading strand and exclude the lagging strand. The looser Mcm2/5 interface at the C-tier ring may have allowed the extrusion of the lagging strand to the outside surface during activation of the CMG helicase. Another major change is the insertion of the Mcm5 C-terminal winged helix domain (WHD) into the interior channel in the active helicase, while the Mcm5 WHD of the inactive Mcm2-7 hexamer is visualized as weak density outside the central channel (Li et al. 2015; Yuan et al. 2016). The Mcm6 C-terminal WHD in the active and inactive Mcm2-7 was also reconfigured. Considering that the WHD is usually involved in DNA binding or interaction with other proteins, the observed movement of these WHDs may have important consequences for helicase activation and replisome function during fork progression.
10.4 GINS Is a Protein-Binding Hub of the Replisome but Does Not Contain a Central Channel for ssDNA
The yeast CMG structure reveals for the first time the complete structure of the GINS heterotetramer (Yuan et al. 2016). The GINS subunits are related in evolution and are composed of two domains each (A and B) (Chang et al. 2007; Choi et al. 2007; Kamada et al. 2007). The B domain of the Psf1 subunit was missing in all previously reported GINS structures (Chang et al. 2007; Choi et al. 2007; Kamada et al. 2007). The Psf1 B domain interacts with Cdc45 by forming an interfacial cross-molecule β-sheet (Yuan et al. 2016). Perhaps the Psf1 B domain is disordered in the absence of Cdc45, explaining why it was not resolved in the crystal structures of the GINS alone.
Earlier low-resolution EM observations as well as the crystal structure of the human GINS suggested a potential hole in the middle of GINS (Boskovic et al. 2007; Chang et al. 2007; Kubota et al. 2003). It was further suggested that ssDNA might thread through the central hole of the GINS. The structure of the full-length yeast GINS contains only a small hole of ~7 Å diameter, too small to thread ssDNA (Fig. 10.3) (Yuan et al. 2016). Importantly, although the GINS structures are very similar between human and yeast, structural features surrounding the central region vary (Chang et al. 2007; Choi et al. 2007; Kamada et al. 2007; Yuan et al. 2016). The yeast GINS has two insertions in Psf1 and Psf2 that are absent in human GINS, and the human GINS has a Psf3 insertion loop that is absent in the yeast GINS. The structural variation, together with the small size, argues against ssDNA threading through the central hole. Instead, the GINS functions mainly to scaffold the replisome by recruiting other proteins. The CMG structure appears to bear this out. Thus GINS interacts extensively with Mcm3/5 and Cdc45 (Yuan et al. 2016). The Psf1 B domain forms the main connection to Cdc45, and the Sld5 subunit binds the Ctf4 trimer scaffold protein that in turn binds to Pol α-primase (Simon et al. 2014). An interaction between Psf1 and the B subunit of Pol ε has also been documented by biochemical studies (Gambus et al. 2009), and recently, the direct interaction between GINS and Pol ε was visualized by negative stain EM (Sun et al. 2015).
10.5 Cdc45 Has Two RecJ-Like Domains and a Protruding Helical Motif
Cdc45 has been predicted to have a RecJ-like fold (Krastanova et al. 2012; Onesti and MacNeill 2013; Sanchez-Pulido and Ponting 2011). The structure of yeast and human Cdc45 confirmed this prediction and further revealed that there are actually two RecJ-like α-/β-domains that are separated by a small helical inter-domain (ID) (Fig. 10.4a–c) (Simon et al. 2016; Yuan et al. 2016). Human and yeast Cdc45 are highly conserved at the structure level and superimpose with a root-mean-square deviation of <2 Å (Fig. 10.4c). Cdc45 has a protruding helical motif (PHM) that contains a highly negatively charged and disordered loop (D166 – R217 in yeast). The protrusion may interact and support the largely flexible N-terminal catalytic domain of Pol2 (see below), because a lysine at the tip of the PHM helix (K222) cross-links to the N-terminal catalytic domain of Pol2 (Yuan et al. 2016). The helical ID contacts and stabilizes the NTDs of Mcm2 and Mcm5 (Fig. 10.1a, c).
The fact that Cdc45 faces the neighboring Mcm2 and Mcm5 subunits in the CMG helicase is suggested to poise Cdc45 to capture DNA if the Mcm2/5 gate breaches open (Petojevic et al. 2015). Although Cdc45 does not have a nuclease activity, there is evidence that Cdc45 binds DNA weakly and nonspecifically (Krastanova et al. 2012; Simon et al. 2016; Bruck and Kaplan 2013). RecJ forms an O-like structure with a pore in the middle to serve as a ssDNA-binding groove, but unlike RecJ the yeast and human Cdc45 do not have a central pore. Therefore, Cdc45 does not contain an internal DNA-binding groove. It is possible that the exterior surface of Cdc45 may help to coordinate DNA in the context of CMG, but the nature of the DNA contact is currently unknown.
Meier-Gorlin syndrome (MGS) is a rare autosomal recessive disorder characterized by short stature (dwarfism) (Klingseisen and Jackson 2011). Previous work has found that mutations in genes involved in establishing the pre-replication complex, such as ORC1, ORC4, ORC6, CDT1, and CDC6, can cause MGS (Hossain and Stillman 2016). More recently, eight point mutations in CDC45 have been identified in MGS patients (Fenwick et al. 2016). The mutations are scattered across the protein structure (Fig. 10.4d). Because Cdc45 is involved in both replication initiation and elongation, it is unclear whether initiation or ongoing fork movement, or both, is the underlying cause of the disease.
10.6 CMG Helicase May Function as an Oil Rig Pump Jack to Inchworm Along DNA
Cryo-EM analysis of CMG particles embedded in vitreous ice revealed two conformations, an extended structure (conformer I) and a compact structure (conformer II) at 4.7 Å and 4.8 Å resolution, respectively (Fig. 10.5a, b) (Yuan et al. 2016). In conformer I, when viewed from the Cdc45-GINS side of the structure, the CTD ring of Mcm2-7 containing the AAA+ motors is tilted by ~10° with respect to the NTD ring, leading to an approximate spiral arrangement of motor domains comprising the CTD ring. In conformer II, the CTD motor ring is approximately parallel to the NTD ring, and CMG is more compact than conformer I. In both conformers, the Mcm2-7 NTD ring-GINS-Cdc45 unit appears to be a rigid platform upon which the CTD AAA+ motor domains switch between extended and compact states during cycles of ATP hydrolysis. These structures suggest that CMG may function like an oil rig pump jack attached to a stable platform, nodding up and down to inchworm along ssDNA and unwind dsDNA at a forked junction (Fig. 10.5c) (Yuan et al. 2016). This linear inchworm ratchet mechanism is distinct from the sequential rotary ATP hydrolysis unwinding mechanism proposed for the homo-hexameric helicases such as E. coli DnaB, E. coli Rho transcription terminator, and bovine papillomavirus (BPV) E1 (Enemark and Joshua-Tor 2006; Itsathitphaisarn et al. 2012; Thomsen and Berger 2009). Additionally, single point mutants of the ATP sites in the Mcm2-7 subunits of Drosophila CMG show major helicase defects in only two of the six mutants, inconsistent with a sequential rotary hydrolysis mechanism that predicts each ATP site would be important to helicase activity (Ilves et al. 2010). Similar conclusions of nonequivalent ATP sites have been made by mutational studies of yeast Mcm2-7 (Schwacha and Bell 2001).
Aside from its main function in unwinding parental DNA, CMG also functions as a scaffold for assembly of the replisome. Indeed, the CMG helicase interacts with many other replicative factors, chief among them are the leading strand Pol ε (Langston et al. 2014) and the Pol α-primase that is required repeatedly on the lagging strand (Gambus et al. 2006).
10.7 Pol ε Binds to the CTD Motor Side of CMG
In 1990 Pol ε was identified as a third DNA polymerase in budding yeast that is essential for cellular replication (Morrison et al. 1990). Subsequent studies revealed that Pol ε is also involved in many other pathways such as the DNA damage checkpoint response, epigenetic silencing, sister chromatid cohesion, and possibly DNA recombination during repair of DNA lesions (Pursell and Kunkel 2008). Nearly two decades after its discovery, Pol ε was assigned as the major leading strand polymerase (Clausen et al. 2015; Kunkel and Burgers 2008; Miyabe et al. 2011; Nick McElhinny et al. 2008; Pursell et al. 2007). In all eukaryotes studied to date, Pol ε consists of four proteins; the largest is the DNA polymerase, followed by the B subunit and two small histone-fold subunits (Pursell and Kunkel 2008). The Pol ε of Saccharomyces cerevisiae consists of Pol2, Dpb2 (DNA polymerase-binding protein 2), Dpb3, and Dpb4. Pol2 is composed of two subdomains: a N-terminal catalytic domain that contains the polymerase and exonuclease active sites and an inactive C-terminal domain that shares homology to B-family DNA polymerases (Fig. 10.6a) (Tahirov et al. 2009). The ternary crystal structure of the N-terminal catalytic subdomain of Pol2 in complex with primer-template DNA and incoming dNTP was recently determined, and it shares many features of other B-family DNA polymerases such as Pol δ (Hogg et al. 2014). However, a P domain, new to the B-family polymerases, enables Pol ε to encircle the nascent dsDNA and enhances the processivity of DNA synthesis. This may explain why Pol ε has the highest fidelity among B-family polymerases despite the absence of an extended β-hairpin loop that is required for high-fidelity replication by other B-family polymerases (Fortune et al. 2005; Hogg and Johansson 2012; Hogg et al. 2014).
Cryo-EM analysis of yeast Pol ε revealed a bilobed overall architecture (Asturias et al. 2006). Pol ε has also recently been shown to bind directly to CMG helicase, independent of DNA, and the 15 protein complex is referred to as CMGE (Langston et al. 2014). Negative stain EM analysis of CMGE provided a 16 Å 3D map (Fig. 10.6b) (Sun et al. 2015). This low-resolution structure revealed that Pol ε is positioned on the CTD-tier side of CMG, sitting atop the CTDs of Mcm2 and Mcm5, and GINS and Cdc45. This Pol ε position was confirmed by extensive cross-linking mass spectrometry analysis (Fig. 10.6c). The visible density of Pol ε could only account for about 70 % the mass of Pol ε, and one suggestion for this missing density was that the catalytic Pol ε N-terminal subdomain was too mobile to visualize (Fig. S6 in ref. (Sun et al. 2015)). However, cross-linking mass spectrometry analysis of CMGE demonstrated that the Pol ε N-terminal subdomain was in proximity to the other subunits of Pol ε and thus located nearby.
10.8 Pol α and Ctf4 Are Located on the NTD-Tier Side of CMG, Opposite from Pol ε
DNA synthesis activity of Pol ε and Pol δ requires the primed site synthesized by the four-subunit Pol α-primase, which synthesizes a hybrid RNA/DNA primer (Conaway and Lehman 1982). A 7-9mer RNA is synthesized de novo by the Pri1/2 subunits followed by an internal transfer of the RNA primer to the DNA polymerase which adds a further 10–20 deoxyribonucleotides (Nethanel et al. 1992). Thus Pol α-primase “counts” the length of RNA and DNA and stops synthesis. The underlying mechanism of how this enzyme functions has been revealed by a series of elegant crystal structures and biochemical studies (Coloma et al. 2016; Kilkenny et al. 2013; Klinge et al. 2009; Perera et al. 2013).
Crystal structures of the yeast Pol α catalytic core (349–1258; 910 amino acids) in the apo form, bound to an RNA/DNA primer template (the binary complex), and in a ternary complex with RNA/DNA primer template and incoming dGTP show that Pol α specifically recognizes the A-form RNA/DNA helix for extension with dNTPs (Perera et al. 2013). But once a full turn of double-helix DNA has been synthesized, Pol α is no longer in direct contact with the A-form RNA/DNA helix, which causes its release from DNA. A recent crystal structure of human Pol α-polymerase subunit in complex with a DNA/DNA helix supports this view, but interestingly, the DNA/DNA helix in contact with the polymerase is in a hybrid A-B form (Coloma et al. 2016). It was suggested that the free energy cost of distorting DNA from B- to the A-B hybrid, rather than the loss of specific contacts, drives the termination of primer synthesis.
Pol α-primase is a bilobed structure in which Pol1 binds to the B subunit and Pri1-Pri2 dimer through its CTD, and the CTD of Pol1 is connected to the polymerase region by a flexible stalk (Nunez-Ramirez et al. 2011). Pol α-primase requires CMG for priming activity on both the leading and lagging strands during unwinding of the forked DNA in the presence of RPA (Georgescu et al. 2015b). Pol α-primase is anchored to CMG helicase by the Ctf4 homotrimer (Gambus et al. 2009; Simon et al. 2014; Tanaka et al. 2009). The C-terminal domain of Ctf4 self-associates into a disk-shaped trimer that binds to a peptide sequence in Sld5, a subunit of the GINS complex within CMG (Simon et al. 2014). The same region of Ctf4 also binds to a homologous peptide sequence within the catalytic subunit of Pol α (Simon et al. 2014). Therefore, Ctf4 acts as a scaffold that cross-links CMG and Pol α-primase, underlying the mechanism by which Ctf4 recruits Pol α-primase to the replisome (Gambus et al. 2009). These physical interactions were visualized in a recent EM analysis of combinations of CMG with Pol ε, Pol α-primase, and Ctf4 (Fig. 10.7) (Sun et al. 2015). Unexpectedly, Pol α-primase and Ctf4 were found to bind to the NTD-tier of CMG, opposite from the CTD-tier where Pol ε binds. This organization was novel because all previous illustrations of the replisome placed the polymerases and primase on the same side of the helicase, after DNA unwinding.
10.9 Future Perspectives
We have reviewed here just some of the numerous recent structural advances in the field of replication that have revealed the structures of replicative polymerases, the 11 subunit CMG helicase, and the overall architecture of the eukaryotic replisome core. There are numerous important questions that remain to be answered, in fact too many to list. One important question for future studies is to understand the DNA path through the replisome and how CMG translocates along DNA. Available biochemical data indicate that CMG functions by steric exclusion in which both the CTD and NTD rings of Mcm2-7 encircle and translocate along one strand of ssDNA, excluding the other strand to the outside of the ring and thus separating the strands (Fu et al. 2011). However, the atomic structure of the Mcm2-7 double hexamer reveals a possible side channel for one DNA strand that lies between the CTD and NTD rings, suggesting the DNA split point may be internal to CMG (Li et al. 2015). Also, the proposed inchworm pump jack mechanism of CMG translocation (Yuan et al. 2016) is a new alternative to the rotary sequential ATP hydrolysis mechanism of hexameric helicase translocation (Enemark and Joshua-Tor 2006; Itsathitphaisarn et al. 2012; Thomsen and Berger 2009). However, there is still very little evidence to support either model. Given the knowledge that CMG helicase moves on the leading strand in a 3′–5′ direction (Bochman and Schwacha 2008; Kang et al. 2012; Moyer et al. 2006), that the 3′ terminus of ssDNA enters CMG through the CTD of Mcm2-7 (Costa et al. 2014), and that Pol ε binds to the CTD side of CMG (Sun et al. 2015), the leading strand DNA may need to travel a long and winding path before reaching Pol ε. These facts, taken together, suggest that the leading strand ssDNA threads through the CTD ring of Mcm2-7, exits the NTD ring of Mcm2-7, and then needs to make a U-turn to reach Pol ε at the CTD side of CMG (Fig. 10.8a). This DNA path assumes that the Mcm2-7 CTD motor ring translocates ahead of the NTD ring and thus pushes against the dsDNA/forked junction, as indicated by EM studies using a biotinylated DNA-streptavidin bound to Drosophila CMG (Costa et al. 2014). Furthermore, in the steric exclusion model of unwinding, the DNA split point would be just before DNA entry into CMG, and thus the newly unwound lagging strand ssDNA would need to traverse the outside surface of the Mcm2-7 ring to reach Pol α-primase on the NTD-tier side of CMG. Why nature would thread DNA through the replisome in this fashion is not understood, considering the extensive ssDNA exposed in this orientation. However, one may propose that this orientation enables Pol ε to be the first to interact with nucleosomes, as Pol ε is known to interact with nucleosomes (Foltman et al. 2013; Tackett et al. 2005) and to function in heterochromatin maintenance (Iida and Araki 2004; Tackett et al. 2005). On the other hand, it is interesting to note that BPV E1 helicase translocates on DNA such that the NTD ring faces the forked junction (Enemark and Joshua-Tor 2008; Lee et al. 2014). If the NTD ring of Mcm2-7 within CMG were to push against the forked junction, this orientation would place Pol α-primase at the top of the fork adjacent to the DNA split point, where Pol α-primase could prime the DNA template immediately after strand separation, and Pol ε would be behind CMG to extend the leading strand immediately after the DNA exits from the Mcm2-7 channel (Fig. 10.8b). We note that CMG helicase and the replisome architectures determined so far are based on in vitro assembled complexes on artificial substrates. It remains to be investigated if the in vivo assembled structures starting from origin DNA are the same especially in the threading of the DNA.
The DNA replication machinery must perform many additional tasks beyond DNA synthesis. For example, the replisome must deal with DNA lesions, transcribing RNA polymerase, DNA-bound proteins, recombination intermediates, nucleosomes in both euchromatin and heterochromatin, and cohesion rings that hold the sister chromosomes together after replication. Thus it may come as no surprise that there are numerous additional proteins that travel with or transiently interact with the replisome, and it is very likely that many new replisome interactive proteins will be identified in future studies. A first glimpse at the multitude of proteins that bind the replisome is provided by pull-outs using antibodies directed against CMG components followed by mass spectrometry to identify CMG-binding proteins (Gambus et al. 2006). This methodology identified a group of proteins referred to as the replisome progression complex (RPC) that contains CMG along with Ctf4, Mcm10, Tof1, Csm3, checkpoint mediator Mrc1, histone chaperone FACT (Spt16 and Pob3), and topoisomerase I (Gambus et al. 2006). The function of FACT in binding H2A/H2B and stimulating transcription through chromatin and of Topo I in relief of supercoiling is understood (Belotserkovskaya et al. 2003; Brill et al. 1987; Kim and Wang 1989; Orphanides et al. 1998; Orphanides et al. 1999; Schlesinger and Formosa 2000). However, most proteins of the RPC have nearly unknown functions. The Ctf4 trimer recognizes a peptide sequence to bind Pol α and Sld5 suggesting that Ctf4 acts as a platform for recruitment and exchange of different replication factors (Villa et al. 2016), in loose analogy to the many proteins that traffic on and off PCNA (Georgescu et al. 2015a). The Mrc1 checkpoint mediator is a stably attached replisome component that appears to aid normal replisome progression in addition to its checkpoint function (Lou et al. 2008), but how Mrc1 facilitates these activities and the spatial orientation of Mrc1 within the replisome are unknown. Tof1 and Csm3 are two proteins that ride with the replisome and appear to pause replication forks at protein-DNA barriers, but how they carry out this function is not understood (Calzada et al. 2005; Mohanty et al. 2006; Tourriere et al. 2005). Mcm10 is an essential replication factor that is reported to interact with ssDNA, dsDNA, Pol α-primase, the CMG helicase, and other factors (reviewed in Bielinsky 2016; Thu and Bielinsky 2013). Mcm10 appears to be the last initiation protein to act at an origin, because two complete CMGs form at an origin yet cannot produce ssDNA until Mcm10 is present (Heller et al. 2011; Kanke et al. 2012; Watase et al. 2012; Yeeles et al. 2015). How Mcm10 fulfills this initiation role and whether Mcm10 moves with the replisome are presently unknown. Besides the RPC proteins, the flap endonuclease 1 (FEN1) and DNA ligase both interact with PCNA and function with Pol δ/PCNA on the lagging strand to remove the RNA primer and join the Okazaki fragments, respectively. How the Pol δ lagging strand polymerase and the RFC clamp loader associate with the replisome, if at all, is another ongoing mystery. Finally, the replisome must replicate through every nucleosome in the genome, including tightly packed heterochromatin without disrupting the epigenetic state of the cell. The CMG, Pol ε, and Pol α all bind nucleosomes or histone subassemblies (Foltman et al. 2013; Huang et al. 2015; Tackett et al. 2005; Wang et al. 2015), yet how the replisome handles nucleosomes and passes them onto the daughter strands is largely unexplored territory. Answers to these many important questions will require a multitude of experimental approaches, but one can be quite certain that cryo-EM structure determination will be essential to understanding the complete picture of these very large and complicated multiprotein assemblies.
Contributor Information
Lin Bai, Cryo-EM Structural Biology Laboratory, Van Andel Research Institute, Grand Rapids, MI, USA.
Zuanning Yuan, Biochemistry and Structural Biology Graduate Program, Stony Brook University, Stony Brook, NY, USA. Cryo-EM Structural Biology Laboratory, Van Andel Research Institute, Grand Rapids, MI, USA.
Jingchuan Sun, Cryo-EM Structural Biology Laboratory, Van Andel Research Institute, Grand Rapids, MI, USA.
Roxana Georgescu, Howard Hughes Medical Institute, The Rockefeller University, New York, NY, USA.
Michael E. O’Donnell, Howard Hughes Medical Institute, The Rockefeller University, New York, NY, USA
Huilin Li, Biochemistry and Structural Biology Graduate Program, Stony Brook University, Stony Brook, NY, USA. Cryo-EM Structural Biology Laboratory, Van Andel Research Institute, Grand Rapids, MI, USA.
References
- Abid Ali F, Renault L, Gannon J, Gahlon HL, Kotecha A, Zhou JC, Rueda D, Costa A. Cryo-EM structures of the eukaryotic replicative helicase bound to a translocation substrate. Nat Commun. 2016;7:10708. doi: 10.1038/ncomms10708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Asturias FJ, Cheung IK, Sabouri N, Chilkova O, Wepplo D, Johansson E. Structure of Saccharomyces cerevisiae DNA polymerase epsilon by cryo-electron microscopy. Nat Struct Mol Biol. 2006;13:35–43. doi: 10.1038/nsmb1040. [DOI] [PubMed] [Google Scholar]
- Banks GR, Boezi JA, Lehman IR. A high molecular weight DNA polymerase from Drosophila melanogaster embryos. Purification, structure, and partial characterization. J Biol Chem. 1979;254:9886–9892. [PubMed] [Google Scholar]
- Bell SD, Botchan MR. The minichromosome maintenance replicative helicase. Cold Spring Harb Perspect Biol. 2013;5:a012807. doi: 10.1101/cshperspect.a012807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bell SP, Labib K. Chromosome duplication in Saccharomyces cerevisiae. Genetics. 2016;203:1027–1067. doi: 10.1534/genetics.115.186452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bell SP, Stillman B. ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex. Nature. 1992;357:128–134. doi: 10.1038/357128a0. [DOI] [PubMed] [Google Scholar]
- Belotserkovskaya R, Oh S, Bondarenko VA, Orphanides G, Studitsky VM, Reinberg D. FACT facilitates transcription-dependent nucleosome alteration. Science. 2003;301:1090–1093. doi: 10.1126/science.1085703. [DOI] [PubMed] [Google Scholar]
- Bielinsky AK. Mcm10: the glue at replication forks. Cell Cycle. 2016;15:1–2. doi: 10.1080/15384101.2016.1216925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bochman ML, Schwacha A. The Mcm2-7 complex has in vitro helicase activity. Mol Cell. 2008;31:287–293. doi: 10.1016/j.molcel.2008.05.020. [DOI] [PubMed] [Google Scholar]
- Bochman ML, Bell SP, Schwacha A. Subunit organization of Mcm2-7 and the unequal role of active sites in ATP hydrolysis and viability. Mol Cell Biol. 2008;28:5865–5873. doi: 10.1128/MCB.00161-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boskovic J, Coloma J, Aparicio T, Zhou M, Robinson CV, Mendez J, Montoya G. Molecular architecture of the human GINS complex. EMBO Rep. 2007;8:678–684. doi: 10.1038/sj.embor.7401002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brewster AS, Wang G, Yu X, Greenleaf WB, Carazo JM, Tjajadi M, Klein MG, Chen XS. Crystal structure of a near-full-length archaeal MCM: functional insights for an AAA+ hexameric helicase. Proc Natl Acad Sci U S A. 2008;105:20191–20196. doi: 10.1073/pnas.0808037105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brill SJ, DiNardo S, Voelkel-Meiman K, Sternglanz R. Need for DNA topoisomerase activity as a swivel for DNA replication for transcription of ribosomal RNA. Nature. 1987;326:414–416. doi: 10.1038/326414a0. [DOI] [PubMed] [Google Scholar]
- Bruck I, Kaplan DL. Cdc45 protein-single-stranded DNA interaction is important for stalling the helicase during replication stress. J Biol Chem. 2013;288:7550–7563. doi: 10.1074/jbc.M112.440941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burgers PM. Polymerase dynamics at the eukaryotic DNA replication fork. J Biol Chem. 2009;284:4041–4045. doi: 10.1074/jbc.R800062200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burgers PM, Gordenin D, Kunkel TA. Who is leading the replication fork, Pol epsilon or Pol delta? Mol Cell. 2016;61:492–493. doi: 10.1016/j.molcel.2016.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calzada A, Hodgson B, Kanemaki M, Bueno A, Labib K. Molecular anatomy and regulation of a stable replisome at a paused eukaryotic DNA replication fork. Genes Dev. 2005;19:1905–1919. doi: 10.1101/gad.337205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang YP, Wang G, Bermudez V, Hurwitz J, Chen XS. Crystal structure of the GINS complex and functional insights into its role in DNA replication. Proc Natl Acad Sci U S A. 2007;104:12685–12690. doi: 10.1073/pnas.0705558104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi JM, Lim HS, Kim JJ, Song OK, Cho Y. Crystal structure of the human GINS complex. Genes Dev. 2007;21:1316–1321. doi: 10.1101/gad.1548107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clausen AR, Lujan SA, Burkholder AB, Orebaugh CD, Williams JS, Clausen MF, Malc EP, Mieczkowski PA, Fargo DC, Smith DJ, et al. Tracking replication enzymology in vivo by genome-wide mapping of ribonucleotide incorporation. Nat Struct Mol Biol. 2015;22:185–191. doi: 10.1038/nsmb.2957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cocker JH, Piatti S, Santocanale C, Nasmyth K, Diffley JF. An essential role for the Cdc6 protein in forming the pre-replicative complexes of budding yeast. Nature. 1996;379:180–182. doi: 10.1038/379180a0. [DOI] [PubMed] [Google Scholar]
- Coloma J, Johnson RE, Prakash L, Prakash S, Aggarwal AK. Human DNA polymerase alpha in binary complex with a DNA:DNA template-primer. Sci Rep. 2016;6:23784. doi: 10.1038/srep23784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Conaway RC, Lehman IR. A DNA primase activity associated with DNA polymerase alpha from Drosophila melanogaster embryos. Proc Natl Acad Sci U S A. 1982;79:2523–2527. doi: 10.1073/pnas.79.8.2523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa A, Ilves I, Tamberg N, Petojevic T, Nogales E, Botchan MR, Berger JM. The structural basis for MCM2-7 helicase activation by GINS and Cdc45. Nat Struct Mol Biol. 2011;18:471–477. doi: 10.1038/nsmb.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa A, Hood IV, Berger JM. Mechanisms for initiating cellular DNA replication. Annu Rev Biochem. 2013;82:25–54. doi: 10.1146/annurev-biochem-052610-094414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa A, Renault L, Swuec P, Petojevic T, Pesavento JJ, Ilves I, MacLellan-Gibson K, Fleck RA, Botchan MR, Berger JM. DNA binding polarity, dimerization, and ATPase ring remodeling in the CMG helicase of the eukaryotic replisome. Elife. 2014;3:e03273. doi: 10.7554/eLife.03273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duzdevich D, Warner MD, Ticau S, Ivica NA, Bell SP, Greene EC. The dynamics of eukaryotic replication initiation: origin specificity, licensing, and firing at the single-molecule level. Mol Cell. 2015;58:483–494. doi: 10.1016/j.molcel.2015.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Enemark EJ, Joshua-Tor L. Mechanism of DNA translocation in a replicative hexameric helicase. Nature. 2006;442:270–275. doi: 10.1038/nature04943. [DOI] [PubMed] [Google Scholar]
- Enemark EJ, Joshua-Tor L. On helicases and other motor proteins. Curr Opin Struct Biol. 2008;18:243–257. doi: 10.1016/j.sbi.2008.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evrin C, Clarke P, Zech J, Lurz R, Sun J, Uhle S, Li H, Stillman B, Speck C. A double-hexameric MCM2-7 complex is loaded onto origin DNA during licensing of eukaryotic DNA replication. Proc Natl Acad Sci U S A. 2009;106:20240–20245. doi: 10.1073/pnas.0911500106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fenwick AL, Kliszczak M, Cooper F, Murray J, Sanchez-Pulido L, Twigg SR, Goriely A, McGowan SJ, Miller KA, Taylor IB, et al. Mutations in CDC45, encoding an essential component of the pre-initiation complex, cause Meier-Gorlin syndrome and Craniosynostosis. Am J Hum Genet. 2016;99:125–138. doi: 10.1016/j.ajhg.2016.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Foltman M, Evrin C, De Piccoli G, Jones RC, Edmondson RD, Katou Y, Nakato R, Shirahige K, Labib K. Eukaryotic replisome components cooperate to process histones during chromosome replication. Cell Rep. 2013;3:892–904. doi: 10.1016/j.celrep.2013.02.028. [DOI] [PubMed] [Google Scholar]
- Forterre P, Filee J, Myllykallio H. Origin and evolution of DNA and DNA replication. Landes Bioscience. 2004:24. [Google Scholar]
- Fortune JM, Pavlov YI, Welch CM, Johansson E, Burgers PM, Kunkel TA. Saccharomyces cerevisiae DNA polymerase delta: high fidelity for base substitutions but lower fidelity for single- and multi-base deletions. J Biol Chem. 2005;280:29980–29987. doi: 10.1074/jbc.M505236200. [DOI] [PubMed] [Google Scholar]
- Froelich CA, Kang S, Epling LB, Bell SP, Enemark EJ. A conserved MCM single-stranded DNA binding element is essential for replication initiation. Elife. 2014;3:e01993. doi: 10.7554/eLife.01993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fu YV, Yardimci H, Long DT, Ho TV, Guainazzi A, Bermudez VP, Hurwitz J, van Oijen A, Scharer OD, Walter JC. Selective bypass of a lagging strand roadblock by the eukaryotic replicative DNA helicase. Cell. 2011;146:931–941. doi: 10.1016/j.cell.2011.07.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gambus A, Jones RC, Sanchez-Diaz A, Kanemaki M, van Deursen F, Edmondson RD, Labib K. GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat Cell Biol. 2006;8:358–366. doi: 10.1038/ncb1382. [DOI] [PubMed] [Google Scholar]
- Gambus A, van Deursen F, Polychronopoulos D, Foltman M, Jones RC, Edmondson RD, Calzada A, Labib K. A key role for Ctf4 in coupling the MCM2-7 helicase to DNA polymerase alpha within the eukaryotic replisome. EMBO J. 2009;28:2992–3004. doi: 10.1038/emboj.2009.226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Georgescu R, Langston L, O’Donnell M. A proposal: evolution of PCNA’s role as a marker of newly replicated DNA. DNA Repair (Amst) 2015a;29:4–15. doi: 10.1016/j.dnarep.2015.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Georgescu RE, Schauer GD, Yao NY, Langston LD, Yurieva O, Zhang D, Finkelstein J, O’Donnell ME. Reconstitution of a eukaryotic replisome reveals suppression mechanisms that define leading/lagging strand operation. Elife. 2015b;4:e04988. doi: 10.7554/eLife.04988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heller RC, Kang S, Lam WM, Chen S, Chan CS, Bell SP. Eukaryotic origin-dependent DNA replication in vitro reveals sequential action of DDK and S-CDK kinases. Cell. 2011;146:80–91. doi: 10.1016/j.cell.2011.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hogg M, Johansson E. DNA polymerase epsilon. Subcell Biochem. 2012;62:237–257. doi: 10.1007/978-94-007-4572-8_13. [DOI] [PubMed] [Google Scholar]
- Hogg M, Osterman P, Bylund GO, Ganai RA, Lundstrom EB, Sauer-Eriksson AE, Johansson E. Structural basis for processive DNA synthesis by yeast DNA polymerase varepsilon. Nat Struct Mol Biol. 2014;21:49–55. doi: 10.1038/nsmb.2712. [DOI] [PubMed] [Google Scholar]
- Hossain M, Stillman B. Meier-Gorlin syndrome. In: Kaplan DL, editor. The initiation of DNA replication in eukaryotes. Springer International Publishing; Cham: 2016. p. 563. [Google Scholar]
- Huang H, Stromme CB, Saredi G, Hodl M, Strandsby A, Gonzalez-Aguilera C, Chen S, Groth A, Patel DJ. A unique binding mode enables MCM2 to chaperone histones H3-H4 at replication forks. Nat Struct Mol Biol. 2015;22:618–626. doi: 10.1038/nsmb.3055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iida T, Araki H. Noncompetitive counteractions of DNA polymerase epsilon and ISW2/yCHRAC for epigenetic inheritance of telomere position effect in Saccharomyces cerevisiae. Mol Cell Biol. 2004;24:217–227. doi: 10.1128/MCB.24.1.217-227.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ilves I, Petojevic T, Pesavento JJ, Botchan MR. Activation of the MCM2-7 helicase by association with Cdc45 and GINS proteins. Mol Cell. 2010;37:247–258. doi: 10.1016/j.molcel.2009.12.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Itsathitphaisarn O, Wing RA, Eliason WK, Wang J, Steitz TA. The hexameric helicase DnaB adopts a nonplanar conformation during translocation. Cell. 2012;151:267–277. doi: 10.1016/j.cell.2012.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson RE, Klassen R, Prakash L, Prakash S. A major role of DNA polymerase delta in replication of both the leading and lagging DNA strands. Mol Cell. 2015;59:163–175. doi: 10.1016/j.molcel.2015.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaguni LS, Rossignol JM, Conaway RC, Banks GR, Lehman IR. Association of DNA primase with the beta/gamma subunits of DNA polymerase alpha from Drosophila melanogaster embryos. J Biol Chem. 1983a;258:9037–9039. [PubMed] [Google Scholar]
- Kaguni LS, Rossignol JM, Conaway RC, Lehman IR. Isolation of an intact DNA polymerase-primase from embryos of Drosophila melanogaster. Proc Natl Acad Sci U S A. 1983b;80:2221–2225. doi: 10.1073/pnas.80.8.2221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kamada K, Kubota Y, Arata T, Shindo Y, Hanaoka F. Structure of the human GINS complex and its assembly and functional interface in replication initiation. Nat Struct Mol Biol. 2007;14:388–396. doi: 10.1038/nsmb1231. [DOI] [PubMed] [Google Scholar]
- Kang YH, Galal WC, Farina A, Tappin I, Hurwitz J. Properties of the human Cdc45/Mcm2-7/GINS helicase complex and its action with DNA polymerase epsilon in rolling circle DNA synthesis. Proc Natl Acad Sci U S A. 2012;109:6042–6047. doi: 10.1073/pnas.1203734109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanke M, Kodama Y, Takahashi TS, Nakagawa T, Masukata H. Mcm10 plays an essential role in origin DNA unwinding after loading of the CMG components. EMBO J. 2012;31:2182–2194. doi: 10.1038/emboj.2012.68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kilkenny ML, Longo MA, Perera RL, Pellegrini L. Structures of human primase reveal design of nucleotide elongation site and mode of Pol alpha tethering. Proc Natl Acad Sci U S A. 2013;110:15961–15966. doi: 10.1073/pnas.1311185110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim RA, Wang JC. Function of DNA topoisomerases as replication swivels in Saccharomyces cerevisiae. J Mol Biol. 1989;208:257–267. doi: 10.1016/0022-2836(89)90387-2. [DOI] [PubMed] [Google Scholar]
- Klinge S, Nunez-Ramirez R, Llorca O, Pellegrini L. 3D architecture of DNA Pol alpha reveals the functional core of multi-subunit replicative polymerases. EMBO J. 2009;28:1978–1987. doi: 10.1038/emboj.2009.150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klingseisen A, Jackson AP. Mechanisms and pathways of growth failure in primordial dwarfism. Genes Dev. 2011;25:2011–2024. doi: 10.1101/gad.169037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krastanova I, Sannino V, Amenitsch H, Gileadi O, Pisani FM, Onesti S. Structural and functional insights into the DNA replication factor Cdc45 reveal an evolutionary relationship to the DHH family of phosphoesterases. J Biol Chem. 2012;287:4121–4128. doi: 10.1074/jbc.M111.285395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kubota Y, Takase Y, Komori Y, Hashimoto Y, Arata T, Kamimura Y, Araki H, Takisawa H. A novel ring-like complex of Xenopus proteins essential for the initiation of DNA replication. Genes Dev. 2003;17:1141–1152. doi: 10.1101/gad.1070003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kunkel TA, Burgers PM. Dividing the workload at a eukaryotic replication fork. Trends Cell Biol. 2008;18:521–527. doi: 10.1016/j.tcb.2008.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langston LD, Zhang D, Yurieva O, Georgescu RE, Finkelstein J, Yao NY, Indiani C, O’Donnell ME. CMG helicase and DNA polymerase epsilon form a functional 15-subunit holoenzyme for eukaryotic leading-strand DNA replication. Proc Natl Acad Sci U S A. 2014;111:15390–15395. doi: 10.1073/pnas.1418334111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee SJ, Syed S, Enemark EJ, Schuck S, Stenlund A, Ha T, Joshua-Tor L. Dynamic look at DNA unwinding by a replicative helicase. Proc Natl Acad Sci U S A. 2014;111:E827–E835. doi: 10.1073/pnas.1322254111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leipe DD, Aravind L, Koonin EV. Did DNA replication evolve twice independently? Nucleic Acids Res. 1999;27:3389–3401. doi: 10.1093/nar/27.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leman AR, Noguchi E. The replication fork: understanding the eukaryotic replication machinery and the challenges to genome duplication. Genes (Basel) 2013;4:1–32. doi: 10.3390/genes4010001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Araki H. Loading and activation of DNA replicative helicases: the key step of initiation of DNA replication. Genes Cells. 2013;18:266–277. doi: 10.1111/gtc.12040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li N, Zhai Y, Zhang Y, Li W, Yang M, Lei J, Tye BK, Gao N. Structure of the eukaryotic MCM complex at 3.8 A. Nature. 2015;524:186–191. doi: 10.1038/nature14685. [DOI] [PubMed] [Google Scholar]
- Liang C, Weinreich M, Stillman B. ORC and Cdc6p interact and determine the frequency of initiation of DNA replication in the genome. Cell. 1995;81:667–676. doi: 10.1016/0092-8674(95)90528-6. [DOI] [PubMed] [Google Scholar]
- Lou H, Komata M, Katou Y, Guan Z, Reis CC, Budd M, Shirahige K, Campbell JL. Mrc1 and DNA polymerase epsilon function together in linking DNA replication and the S phase checkpoint. Mol Cell. 2008;32:106–117. doi: 10.1016/j.molcel.2008.08.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- MacNeill S. Composition and dynamics of the eukaryotic replisome: a brief overview. Subcell Biochem. 2012;62:1–17. doi: 10.1007/978-94-007-4572-8_1. [DOI] [PubMed] [Google Scholar]
- Meselson M, Stahl FW. The replication of DNA in Escherichia Coli. Proc Natl Acad Sci U S A. 1958;44:671–682. doi: 10.1073/pnas.44.7.671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller JM, Arachea BT, Epling LB, Enemark EJ. Analysis of the crystal structure of an active MCM hexamer. Elife. 2014;3:e03433. doi: 10.7554/eLife.03433. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mimura S, Seki T, Tanaka S, Diffley JF. Phosphorylation-dependent binding of mitotic cyclins to Cdc6 contributes to DNA replication control. Nature. 2004;431:1118–1123. doi: 10.1038/nature03024. [DOI] [PubMed] [Google Scholar]
- Miyabe I, Kunkel TA, Carr AM. The major roles of DNA polymerases epsilon and delta at the eukaryotic replication fork are evolutionarily conserved. PLoS Genet. 2011;7:e1002407. doi: 10.1371/journal.pgen.1002407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohanty BK, Bairwa NK, Bastia D. The Tof1p-Csm3p protein complex counteracts the Rrm3p helicase to control replication termination of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2006;103:897–902. doi: 10.1073/pnas.0506540103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morrison A, Araki H, Clark AB, Hamatake RK, Sugino A. A third essential DNA polymerase in S. cerevisiae. Cell. 1990;62:1143–1151. doi: 10.1016/0092-8674(90)90391-q. [DOI] [PubMed] [Google Scholar]
- Moyer SE, Lewis PW, Botchan MR. 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. 2006;103:10236–10241. doi: 10.1073/pnas.0602400103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nethanel T, Zlotkin T, Kaufmann G. Assembly of simian virus 40 Okazaki pieces from DNA primers is reversibly arrested by ATP depletion. J Virol. 1992;66:6634–6640. doi: 10.1128/jvi.66.11.6634-6640.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nick McElhinny SA, Gordenin DA, Stith CM, Burgers PM, Kunkel TA. Division of labor at the eukaryotic replication fork. Mol Cell. 2008;30:137–144. doi: 10.1016/j.molcel.2008.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nishitani H, Lygerou Z, Nishimoto T, Nurse P. The Cdt1 protein is required to license DNA for replication in fission yeast. Nature. 2000;404:625–628. doi: 10.1038/35007110. [DOI] [PubMed] [Google Scholar]
- Nunez-Ramirez R, Klinge S, Sauguet L, Melero R, Recuero-Checa MA, Kilkenny M, Perera RL, Garcia-Alvarez B, Hall RJ, Nogales E, et al. Flexible tethering of primase and DNA Pol alpha in the eukaryotic primosome. Nucleic Acids Res. 2011;39:8187–8199. doi: 10.1093/nar/gkr534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- O’Donnell M, Langston L, Stillman B. Principles and concepts of DNA replication in bacteria, archaea, and eukarya. Cold Spring Harb Perspect Biol. 2013:5. doi: 10.1101/cshperspect.a010108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Onesti S, MacNeill SA. Structure and evolutionary origins of the CMG complex. Chromosoma. 2013;122:47–53. doi: 10.1007/s00412-013-0397-x. [DOI] [PubMed] [Google Scholar]
- Orphanides G, LeRoy G, Chang CH, Luse DS, Reinberg D. FACT, a factor that facilitates transcript elongation through nucleosomes. Cell. 1998;92:105–116. doi: 10.1016/s0092-8674(00)80903-4. [DOI] [PubMed] [Google Scholar]
- Orphanides G, Wu WH, Lane WS, Hampsey M, Reinberg D. The chromatin-specific transcription elongation factor FACT comprises human SPT16 and SSRP1 proteins. Nature. 1999;400:284–288. doi: 10.1038/22350. [DOI] [PubMed] [Google Scholar]
- Perera RL, Torella R, Klinge S, Kilkenny ML, Maman JD, Pellegrini L. Mechanism for priming DNA synthesis by yeast DNA polymerase alpha. Elife. 2013;2:e00482. doi: 10.7554/eLife.00482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Petojevic T, Pesavento JJ, Costa A, Liang J, Wang Z, Berger JM, Botchan MR. Cdc45 (cell division cycle protein 45) guards the gate of the Eukaryote replisome helicase stabilizing leading strand engagement. Proc Natl Acad Sci U S A. 2015;112:E249–E258. doi: 10.1073/pnas.1422003112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pursell ZF, Kunkel TA. DNA polymerase epsilon: a polymerase of unusual size (and complexity) Prog Nucleic Acid Res Mol Biol. 2008;82:101–145. doi: 10.1016/S0079-6603(08)00004-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pursell ZF, Isoz I, Lundstrom EB, Johansson E, Kunkel TA. Yeast DNA polymerase epsilon participates in leading-strand DNA replication. Science. 2007;317:127–130. doi: 10.1126/science.1144067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Remus D, Beuron F, Tolun G, Griffith JD, Morris EP, Diffley JF. Concerted loading of Mcm2-7 double hexamers around DNA during DNA replication origin licensing. Cell. 2009;139:719–730. doi: 10.1016/j.cell.2009.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sanchez-Pulido L, Ponting CP. Cdc45: the missing RecJ ortholog in eukaryotes? Bioinformatics. 2011;27:1885–1888. doi: 10.1093/bioinformatics/btr332. [DOI] [PubMed] [Google Scholar]
- Santocanale C, Diffley JF. ORC- and Cdc6-dependent complexes at active and inactive chromosomal replication origins in Saccharomyces cerevisiae. EMBO J. 1996;15:6671–6679. [PMC free article] [PubMed] [Google Scholar]
- Schlesinger MB, Formosa T. POB3 is required for both transcription and replication in the yeast Saccharomyces cerevisiae. Genetics. 2000;155:1593–1606. doi: 10.1093/genetics/155.4.1593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwacha A, Bell SP. Interactions between two catalytically distinct MCM subgroups are essential for coordinated ATP hydrolysis and DNA replication. Mol Cell. 2001;8:1093–1104. doi: 10.1016/s1097-2765(01)00389-6. [DOI] [PubMed] [Google Scholar]
- Simon AC, Zhou JC, Perera RL, van Deursen F, Evrin C, Ivanova ME, Kilkenny ML, Renault L, Kjaer S, Matak-Vinkovic D, et al. A Ctf4 trimer couples the CMG helicase to DNA polymerase alpha in the eukaryotic replisome. Nature. 2014;510:293–297. doi: 10.1038/nature13234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simon AC, Sannino V, Costanzo V, Pellegrini L. Structure of human Cdc45 and implications for CMG helicase function. Nat Commun. 2016;7:11638. doi: 10.1038/ncomms11638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singleton MR, Sawaya MR, Ellenberger T, Wigley DB. Crystal structure of T7 gene 4 ring helicase indicates a mechanism for sequential hydrolysis of nucleotides. Cell. 2000;101:589–600. doi: 10.1016/s0092-8674(00)80871-5. [DOI] [PubMed] [Google Scholar]
- Slaymaker IM, Chen XS. MCM structure and mechanics: what we have learned from archaeal MCM. Subcell Biochem. 2012;62:89–111. doi: 10.1007/978-94-007-4572-8_6. [DOI] [PubMed] [Google Scholar]
- Steitz TA. DNA polymerases: structural diversity and common mechanisms. J Biol Chem. 1999;274:17395–17398. doi: 10.1074/jbc.274.25.17395. [DOI] [PubMed] [Google Scholar]
- Sun J, Evrin C, Samel SA, Fernandez-Cid A, Riera A, Kawakami H, Stillman B, Speck C, Li H. Cryo-EM structure of a helicase loading intermediate containing ORC-Cdc6-Cdt1-MCM2-7 bound to DNA. Nat Struct Mol Biol. 2013;20:944–951. doi: 10.1038/nsmb.2629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun J, Fernandez-Cid A, Riera A, Tognetti S, Yuan Z, Stillman B, Speck C, Li H. Structural and mechanistic insights into Mcm2-7 double-hexamer assembly and function. Genes Dev. 2014;28:2291–2303. doi: 10.1101/gad.242313.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun J, Shi Y, Georgescu RE, Yuan Z, Chait BT, Li H, O’Donnell ME. The architecture of a eukaryotic replisome. Nat Struct Mol Biol. 2015;22:976–982. doi: 10.1038/nsmb.3113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tackett AJ, Dilworth DJ, Davey MJ, O’Donnell M, Aitchison JD, Rout MP, Chait BT. Proteomic and genomic characterization of chromatin complexes at a boundary. J Cell Biol. 2005;169:35–47. doi: 10.1083/jcb.200502104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tahirov TH, Makarova KS, Rogozin IB, Pavlov YI, Koonin EV. 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. 2009;4:11. doi: 10.1186/1745-6150-4-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanaka S, Araki H. Helicase activation and establishment of replication forks at chromosomal origins of replication. Cold Spring Harb Perspect Biol. 2013;5:a010371. doi: 10.1101/cshperspect.a010371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanaka S, Diffley JF. Interdependent nuclear accumulation of budding yeast Cdt1 and Mcm2-7 during G1 phase. Nat Cell Biol. 2002;4:198–207. doi: 10.1038/ncb757. [DOI] [PubMed] [Google Scholar]
- Tanaka H, Katou Y, Yagura M, Saitoh K, Itoh T, Araki H, Bando M, Shirahige K. Ctf4 coordinates the progression of helicase and DNA polymerase alpha. Genes Cells. 2009;14:807–820. doi: 10.1111/j.1365-2443.2009.01310.x. [DOI] [PubMed] [Google Scholar]
- Thomsen ND, Berger JM. Running in reverse: the structural basis for translocation polarity in hexameric helicases. Cell. 2009;139:523–534. doi: 10.1016/j.cell.2009.08.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thu YM, Bielinsky AK. Enigmatic roles of Mcm10 in DNA replication. Trends Biochem Sci. 2013;38:184–194. doi: 10.1016/j.tibs.2012.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ticau S, Friedman LJ, Ivica NA, Gelles J, Bell SP. Single-molecule studies of origin licensing reveal mechanisms ensuring bidirectional helicase loading. Cell. 2015;161:513–525. doi: 10.1016/j.cell.2015.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tognetti S, Riera A, Speck C. Switch on the engine: how the eukaryotic replicative helicase MCM2-7 becomes activated. Chromosoma. 2015;124:13–26. doi: 10.1007/s00412-014-0489-2. [DOI] [PubMed] [Google Scholar]
- Tourriere H, Versini G, Cordon-Preciado V, Alabert C, Pasero P. Mrc1 and Tof1 promote replication fork progression and recovery independently of Rad53. Mol Cell. 2005;19:699–706. doi: 10.1016/j.molcel.2005.07.028. [DOI] [PubMed] [Google Scholar]
- Villa F, Simon AC, Ortiz Bazan MA, Kilkenny ML, Wirthensohn D, Wightman M, Matak-Vinkovic D, Pellegrini L, Labib K. Ctf4 is a hub in the eukaryotic replisome that links multiple CIP-box proteins to the CMG helicase. Mol Cell. 2016;63:385–396. doi: 10.1016/j.molcel.2016.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang H, Wang M, Yang N, Xu RM. Structure of the quaternary complex of histone H3–H4 heterodimer with chaperone ASF1 and the replicative helicase subunit MCM2. Protein Cell. 2015;6:693–697. doi: 10.1007/s13238-015-0190-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watase G, Takisawa H, Kanemaki MT. Mcm10 plays a role in functioning of the eukaryotic replicative DNA helicase, Cdc45-Mcm-GINS. Curr Biol. 2012;22:343–349. doi: 10.1016/j.cub.2012.01.023. [DOI] [PubMed] [Google Scholar]
- Watson JD, Crick FH. Genetical implications of the structure of deoxyribonucleic acid. Nature. 1953a;171:964–967. doi: 10.1038/171964b0. [DOI] [PubMed] [Google Scholar]
- Watson JD, Crick FH. Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature. 1953b;171:737–738. doi: 10.1038/171737a0. [DOI] [PubMed] [Google Scholar]
- Yeeles JT, Deegan TD, Janska A, Early A, Diffley JF. Regulated eukaryotic DNA replication origin firing with purified proteins. Nature. 2015;519:431–435. doi: 10.1038/nature14285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuan Z, Bai L, Sun J, Georgescu R, Liu J, O’Donnell ME, Li H. Structure of the eukaryotic replicative CMG helicase suggests a pumpjack motion for translocation. Nat Struct Mol Biol. 2016;23:217–224. doi: 10.1038/nsmb.3170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang D, O’Donnell M. The eukaryotic replication machine. Enzyme. 2016;39:191–229. doi: 10.1016/bs.enz.2016.03.004. [DOI] [PubMed] [Google Scholar]