Abstract
Cellular DNA replication is initiated through the action of multiprotein complexes that recognize replication start sites in the chromosome (termed origins) and facilitate duplex DNA melting within these regions. In a given cell cycle, initiation occurs only once per origin and each round of replication is tightly coupled to cell division. To avoid aberrant origin firing and re-replication, eukaryotes tightly regulate two events in the initiation process: loading of the replicative helicase, MCM2-7, onto chromatin by the Origin Recognition Complex (ORC), and subsequent activation of the helicase by incorporation into a complex known as the CMG. Recent work has begun to reveal the details of an orchestrated and sequential exchange of initiation factors on DNA that give rise to a replication-competent complex, the replisome. Here we review the molecular mechanisms that underpin eukaryotic DNA replication initiation – from selecting replication start sites to replicative helicase loading and activation – and describe how these events are often distinctly regulated across different eukaryotic model organisms.
Keywords: DNA replication, ORC, Cdc6, initiator, Cdt1, MCM2-7, CMG, helicase
Introduction
The success of biological organisms depends on the faithful transmission of genetic information from parent to progeny. All life-forms store their genetic content in the form of nucleic acids, and the replication and dissemination of this information forms the fundamental basis of inheritance. In cells, the process of replication involves two primary tasks: (1) the separation of duplex DNA into two single-stranded templates and (2) semi-conservative replication of each strand. These events are coupled with cell division to produce progeny with essentially identical copies of the parent’s genetic information. Through all cellular lineages, a conserved division of labor has been applied to the process of DNA replication, such that separable tasks (e.g., start site selection, duplex unwinding, DNA synthesis) are allocated to different, albeit sometimes overlapping, factors. Although this basic framework is conserved throughout the bacterial, archaeal, and eukaryote kingdoms, there has been significant evolutionary diversification of the molecules that complete each task, to the point where it is now clear that aspects of the replicative machinery emerged twice, independently, during cellular evolution (Edgell and Doolittle, 1997, Leipe et al., 1999). Replication initiation in eukaryl species has become particularly elaborated by disparate forms of regulation to meet the specific demands of multicellularity, development and large genome size.
Replication is started by a trans-acting “initiator” factor that directs, in both space and time, loading of the replicative machinery onto particular genomic loci known as origins. In general, the number of origins scales with genome size, thereby ensuring that chromosome duplication can be carried out on a physiologically manageable timescale (Gilbert, 2004). Bacteria, as well as certain archaea, frequently initiate replication with a single chromosomal start site (Costa et al., 2013, Wu et al., 2014b). Conversely, some archaeal chromosomes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes (the 12 Mbp S. cerevisiae genome contains around 400 origins and the 3 Gbp human genome ~30,000–50,000 origins (Leonard and Mechali, 2013)). The large number of origins and the need to coordinate initiation across these sites represents a fundamental challenge to DNA replication in eukaryotes; other factors, such as the use of multiple linear chromosomes, as opposed to a single circular chromosome, add additional replicative complexity. Replication is particularly problematic in multicellular eukaryotes, where the process of development can alter replication timing and frequency, and in the context of cellular differentiation, which changes the chromatin landscape and requires the coordinated transmission of epigenetic marks. Despite these and other challenges, more than 15 trillion repetitive rounds of DNA replication and cell division are successfully executed on the developmental path of a fertilized human embryo to the adult human body (Bianconi et al., 2013). Although our understanding of the replicative process is far from complete, we are beginning to understand how eukaryotes utilize a variety of sequential and redundant regulatory mechanisms to achieve this biological feat.
The eukaryotic replisome is built from the regulated and stepwise assembly of multiple intermediary replication factor complexes. In S. cerevisiae, 42 individual proteins are sufficient to fully reconstitute DNA replication in vitro, and since many of these proteins function within large macromolecular assemblies, fewer than fifteen pre-assembled replication factors are required (Yeeles et al., 2015). In short, replication initiation entails four steps (Figure 1): (1) demarcation of start sites by the Origin Recognition Complex (ORC) and the Cdc6 helicase-loader; (2) reiterative loading of an inactive form of the replicative helicase, MCM2-7, by ORC•Cdc6 and the Cdt1 chaperone to form the pre-replication complex (pre-RC); (3) helicase activation by the formation of the Cdc45•MCM2-7•GINS (CMG) complex (the pre-initiation complex, pre-IC); and (4) generation of a bidirectional replication fork that depends on prior origin melting by the MCM2-7 complex and on the tethering of DNA polymerases and additional accessory factors to the replicative helicase. Here we review the molecular mechanisms underpinning eukaryotic replication initiation, from origin specification to helicase activation. This review, although focused on eukaryotic mechanisms, will, as needed, reference studies of the archaeal system to fill critical gaps in knowledge. Given the extensive number of publications in the field, we apologize to those colleagues whose work is not referenced due to space limitations.
Figure 1.
Mechanistic outline of DNA replication initiation in eukarya. During the G1 phase of the cell cycle, an origin-bound ORC•Cdc6 complex together with Cdt1 facilitates the sequential recruitment and loading of two MCM2-7 complexes into a stable double hexamer that encircles duplex DNA (pre-RC). At the onset of S phase, the helicase is activated, leading to origin unwinding. The recruitment of other initiation factors (Cdc45 and GINS, the pre-IC) and double-hexamer dissolution activate the helicase to drive fork progression as a single-stranded DNA-bound Cdc45•MCM2-7•GINS (CMG) complex. A color version of this figure is available online.
Origins of replication
Origins of replication are chromosomal regions that recruit replication initiators for facilitating assembly of the replication machinery (Francois Jacob, 1963). The defining features of eukaryotic origins are complicated and continuously evolving; for more thorough coverage we refer the reader to a number of excellent reviews on the topic (MacAlpine and Bell, 2005, Leonard and Mechali, 2013, Creager et al., 2015). Here we briefly discuss the most salient features of origins in S. cerevisiae, S. pombe, and metazoans, with a particular focus on details pertinent to later topics of discussion.
S. cerevisiae origins of replication
S. cerevisiae origins of replication were initially identified as chromosomal regions capable of conferring replicative properties to exogenous plasmids (Stinchcomb et al., 1979). These AT-rich, autonomous replication sequences (ARSs) function as replication start sites (Huberman et al., 1987, Brewer and Fangman, 1987, Huberman et al., 1988) and serve to recruit the eukaryotic initiator, ORC (Bell and Stillman, 1992). The ARS contains a number of essential elements, the most important of which is the eleven-basepair ‘A element’, which constitutes the ARS consensus sequence (ACS) and represents the primary site of initiator binding (Marahrens and Stillman, 1992, Rao et al., 1994, Theis and Newlon, 1994, Rao and Stillman, 1995). Notably, amongst the different model eukaryotic systems used for studying replication, only S. cerevisiae appears to utilize a specific consensus sequence (Figure 2).
Figure 2.
Molecular details of eukaryotic origins and mechanisms of ORC binding. (GREY) S. cerevisiae origins are distinctive among eukaryotes for conforming to a consensus sequence, the ACS. ScORC can bind the ACS directly and specifically, although interactions between the Orc1-BAH domain and nucleosomes can also modulate ORC origin selection. (PURPLE) Although they do not possess a strict consensus sequence, S. pombe origins are AT-rich. SpORC specifically binds such sites using a domain insertion unique to SpOrc4 that encodes a DNA-binding AT-hook motif. (GREEN) Metazoan ORC can be targeted to chromosomes through a variety of mechanisms, including the Orc1 BAH domain, the DNA-binding TFIIB domain of Orc6, and through interactions with chromatin-associated factors. A majority of metazoan origins are also predicted to contain G-quadruplex secondary structure elements, but how this feature affects ORC binding is currently unclear. A color version of this figure is available online.
Although there are over 12,000 ACS-like sequences in the yeast genome, only about 400 facilitate replication initiation (Wyrick et al., 2001, Nieduszynski et al., 2006, Xu et al., 2006). This low usage of possible S. cerevisiae origins (< 5%) derives in part from an additional level of origin specification that is imposed by the local chromatin structure. S. cerevisiae origins, like those of other eukaryotes, are maintained as nucleosome-free regions (NFRs) (Eaton et al., 2010, Berbenetz et al., 2010). However, surrounding an ARS, ORC-dependent nucleosome phasing directly affects the efficiency of origin usage (Thoma et al., 1984, Simpson, 1990, Lipford and Bell, 2001, Berbenetz et al., 2010), and the eviction of ORC results in nucleosome encroachment into ARS regions (Thoma et al., 1984, Eaton et al., 2010). Mechanistically, NFRs allow ORC access to DNA, while phased nucleosomes provide additional favorable sites for binding ORC (Muller et al., 2010, Hizume et al., 2013). Thus, S. cerevisiae origins are defined by ORC binding to both an ACS and specific chromatin features (Leonard and Mechali, 2013, Hoggard et al., 2013).
S. pombe origins of replication
Genome-wide studies demonstrate that the S. pombe genome contains around 400 origin sequences that are generally nucleosome-free (Givens et al., 2012, Xu et al., 2012), AT-rich, and around 1 kilobase long (Segurado et al., 2003, Dai et al., 2005, Heichinger et al., 2006). These findings are consistent with biochemical studies showing that origin usage in fission yeast depends on clustered stretches of adenine and thymine (Clyne and Kelly, 1995, Kim et al., 2001, Okuno et al., 1999, Dai et al., 2005). Origin selection in S. pombe is facilitated by a species-specific insertion in the Orc4 subunit of ORC that encodes a DNA-binding element that specifically recognizes the minor groove of AT-rich sequences (Figure 2) (Chuang and Kelly, 1999, Moon et al., 1999).
Metazoan origins of replication
Unlike budding yeast ORC, which shows a degree of sequence-specificity, metazoan ORC binds DNA promiscuously (Vashee et al., 2003, Remus et al., 2004). This behavior is consistent with the observation that metazoan replication initiates from diverse sequences (Mechali and Kearsey, 1984, Heinzel et al., 1991, Hyrien and Mechali, 1993). Despite origin sequence variability, the genome-wide analysis of replication start sites has revealed some common patterns in metazoan origins. As with budding and fission yeast, metazoan ORC binds to NFRs in the genome (MacAlpine et al., 2010, Karnani et al., 2010, Eaton et al., 2011), which in turn favorably contributes to assembly of the replication machinery (Lubelsky et al., 2011). Interestingly, G-rich sequences and CpG islands are highly enriched in metazoan origins (Delgado et al., 1998, Cadoret et al., 2008, Prioleau, 2009, Sequeira-Mendes et al., 2009) and have been proposed to serve two purposes: NFR maintenance (Huppert and Balasubramanian, 2007, Wong and Huppert, 2009, Fenouil et al., 2012) and the favoring of G-quadraplex formation (Cayrou et al., 2011, Cayrou et al., 2012, Valton et al., 2014, Cayrou et al., 2015). Preliminary analyses suggest that ORC may preferentially associate with G-rich elements (Zellner et al., 2007, Hoshina et al., 2013), indicating that a conserved structural feature in DNA, rather than a specific consensus sequence, may aid ORC binding in metazoans. As in S. cerevisiae, metazoan origin selection by ORC is further fine-tuned by direct interactions with nucleosomes and chromatin-associated factors. For example, the N-terminal Bromo-Adjacent Homology (BAH) domain in Orc1 directly interacts with histones to direct origin usage (Figure 2) (Noguchi et al., 2006, Muller et al., 2010, Kuo et al., 2012).
Although ORC can be targeted to specific genomic loci, origins cannot be strictly defined by the position of ORC binding on chromatin. Indeed, once loaded onto DNA by ORC, the MCM2-7 helicase (which eventually nucleates replisome assembly following duplex melting) appears to lack positional restraints and is free to either diffuse away from ORC (Remus et al., 2009, Evrin et al., 2009) or be forcibly displaced by other chromatin-localized cellular processes (such as the transcriptional machinery) (Ritzi et al., 1998, Edwards et al., 2002, Powell et al., 2015, Gros et al., 2015). Thus, origins in eukaryotes must be defined flexibly, as the site of initiator binding does not always reflect the site of replication initiation.
Overall, both yeast and metazoa contain conserved sequence elements at origins that are known or proposed to guide ORC binding. In yeast, these elements are encoded within the DNA primary sequence and in metazoa potentially by a propensity to form distinctive secondary structures. Chromatin context plays an additional critical, but poorly understood role in origin usage in all eukaryotes. Understanding how these cis- and trans-acting origin elements interface is an important and active area of future research.
The Origin Recognition Complex (ORC)
Eukaryotic origins direct the recruitment of the Origin Recognition Complex (ORC), a conserved heterohexameric protein assembly identified for its ability to specifically recognize the double-stranded form of the yeast ACS (Bell and Stillman, 1992). Upon recruitment to chromosomal replication start sites, ORC binds an additional factor, Cdc6, as a necessary prerequisite to helicase loading. Despite its centrality to ORC function, the mechanism of DNA binding by the initiator has long remained ambiguous. However, structural studies from archaea and eukaryotes have revealed a conserved mechanism for the association of ORC with DNA that informs not only our understanding of how the ORC•Cdc6 initiator stably binds replication start sites, but also how ORC mediates downstream helicase loading events (Dueber et al., 2007, Gaudier et al., 2007, Sun et al., 2013, Bleichert et al., 2015).
The Origin Recognition Complex and Cdc6
ORC was first identified by fractionation of ARS-binding proteins in budding yeast (Bell and Stillman, 1992). Although many other ARS-binding factors had been identified previously (Jazwinski and Edelman, 1982, Sweder et al., 1988, Diffley and Stillman, 1988, Buchman et al., 1988, Shore et al., 1987), ORC proved uniquely able to bind the ACS (Diffley and Cocker, 1992, Marahrens and Stillman, 1992, Li and Herskowitz, 1993), and temperature sensitive mutants exhibited cell-cycle arrest at a stage consistent with a role in the early aspects of DNA replication (Bell et al., 1993, Foss et al., 1993, Micklem et al., 1993). Following the discovery of ORC in budding yeast, the broad eukaryotic conservation of ORC was demonstrated with the identification of orthologs in S. pombe (Muzi-Falconi and Kelly, 1995, Grallert and Nurse, 1996), D. melanogaster (Gossen et al., 1995, Landis et al., 1997), X. laevis (Carpenter et al., 1996, Romanowski et al., 1996b), and humans (Gavin et al., 1995). The mechanism of ORC function has now been investigated across multiple model organisms, revealing that the core components, subunit organization, and function of ORC are broadly conserved (although not universally, particularly in protozoa (El-Sayed et al., 2005)). Despite this conservation, there do exist certain species-specific alterations that appear to have generated notable functional differences.
S. cerevisiae ORC is a roughly 400 kDa assembly composed of six proteins (Orc1-6) named in descending order of molecular weight (Bell et al., 1993). Orc1-5 exhibit relatively good conservation across species and are members of the ATPases associated with diverse cellular activities (AAA+) superfamily of proteins; Orc1-5 also contain a C-terminal winged-helix (WH) domain (Loo et al., 1995, Bell et al., 1995, Muzi-Falconi and Kelly, 1995, Liu et al., 2000). Overall, the Orc1-5 AAA+ and WH domains, which account for a bit over half the total mass of ORC, display an average of ~45–50% similarity and 25–30% identity between yeast and human orthologs (Tugal et al., 1998, Speck et al., 2005). A majority of AAA+ family members bind and hydrolyze ATP through the formation of a composite, bipartite active site that is generated between neighboring AAA+ protomers upon subunit oligomerization (Guenther et al., 1997, Lenzen et al., 1998, Putnam et al., 2001). AAA+ proteins, like the broad P-loop family of NTPases to which they belong (Iyer et al., 2004), contain so-called ‘Walker A’ and ‘Walker B’ signature sequence motifs that contribute to nucleotide binding and hydrolysis, respectively. A third motif, generally critical to AAA+ ATPase function, is the arginine finger, which is presented in trans to the neighboring subunit to reconstitute an active ATPase site (Neuwald et al., 1999, Zhang et al., 2000). While many of the structural elements important for ATP binding are conserved in Orc1, Orc4, and Orc5, only Orc1 has been found to possess ATPase activity (Klemm et al., 1997, Makise et al., 2003, Ranjan and Gossen, 2006, Siddiqui and Stillman, 2007, Bleichert et al., 2015); Orc2 and Orc3 retain degenerate AAA+ scaffolds that lack functional active site motifs altogether (Speck et al., 2005, Clarey et al., 2006, Bleichert et al., 2015). Intriguingly, although many AAA+ family members function as toroidal hexameric assemblies (reviewed in (Hanson and Whiteheart, 2005)), ORC retains only five proteins with AAA+ domains. Orc6 lacks a AAA+ domain (Chesnokov et al., 2003, Balasov et al., 2007, Bleichert et al., 2013) and its primary sequence is only weakly conserved between yeast and human, making it the least conserved ORC subunit (Dhar and Dutta, 2000). Nonetheless, certain elements of Orc6 are conserved across species, including an N-terminal TFIIB-like domain and a short conserved region at the extreme C-terminus of the protein (Liu et al., 2011, Bleichert et al., 2013).
Cdc6 forms a complex with ORC at origins and is required for initiator function. Cdc6 was first identified in S. cerevisiae mutant screens (Hartwell, 1976) and was later found to have a role in replication initiation (Palmer et al., 1990), with functional requirements prior to S-phase (Kelly et al., 1993, Hogan and Koshland, 1992, Zwerschke et al., 1994). A genetic interaction between ORC and Cdc6 suggested a coordinated activity for these factors (Liang et al., 1995), with supporting data demonstrating an ORC-dependent recruitment of Cdc6 to origins that depends on a direct interaction between the proteins (Santocanale and Diffley, 1996, Leatherwood et al., 1996, Grallert and Nurse, 1996, Coleman et al., 1996, Cocker et al., 1996, Liang et al., 1995, Kong et al., 2003). Importantly, analysis of the Cdc6 primary sequence reveals conserved nucleotide binding and hydrolysis motifs (Lisziewicz et al., 1988, Zhou et al., 1989, Zwerschke et al., 1994), as well as close homology to ORC’s AAA+ subunits, particularly Orc1 (Bell et al., 1995, Quintana et al., 1997). Cdc6 associates with chromatin-bound ORC in an ATP-dependent fashion (Perkins and Diffley, 1998, Kneissl et al., 2003, Evrin et al., 2013, Kang et al., 2014, Coster et al., 2014, Ticau et al., 2015), an interaction that activates Cdc6’s ATPase activity and is consistent with the reconstitution of canonical AAA+ interactions (Randell et al., 2006, Speck and Stillman, 2007). Thus, Cdc6 recruitment provides a sixth AAA+ subunit to the initiator complex overall.
Structure of ORC and ORC•Cdc6
Orc/Cdc6 homologs have been identified in all eukaryotic and archaeal species analyzed (Aves et al., 2012). In many archaea, the initiation factors are genetically streamlined such that certain species possess only a single Orc/Cdc6 gene (Barry and Bell, 2006) and the relative simplicity of the archaeal system has been exploited to help understand the structure and function of eukaryotic Orc homologs. A conserved three-domain architecture is observed for Orc and Cdc6 homologs, with two domains at the N-terminus forming the central AAA+ module and a third at the C-terminus comprising a loosely-tethered WH domain (Figure 3A) (Liu et al., 2000, Singleton et al., 2004, Gaudier et al., 2007, Dueber et al., 2007, Bleichert et al., 2015). The structure of archaeal Orc in complex with DNA has shown that both the WH and AAA+ domains contact DNA (Gaudier et al., 2007, Dueber et al., 2007), and biochemical studies have demonstrated that contacts mediated by the AAA+ domain contribute to origin specificity (Dueber et al., 2011). Surprisingly, structures of Orc/DNA complexes have not revealed evidence for the formation of an ATPase-competent Orc dimer (Gaudier et al., 2007, Dueber et al., 2007), despite the existence of a conserved arginine finger in these proteins that would otherwise suggest that oligomerization might occur. The mechanism that promotes formation of a catalytically active ATPase in the archaeal system is unclear.
Figure 3.
ORC architecture. A) Cdc6/Orc homologs are characterized by three domains, two of which form the AAA+ module (green) and a third that encodes a winged-helix (WH) domain (grey). Bound nucleotide is shown as sticks (PDB = 1FNN). B) The D. melanogaster ORC heterohexamer is a crescent shaped molecule with the AAA+ (green surface) and WH (grey cartoon) domains forming a domain-swapped arrangement. Orc6 is bound by a domain insertion in the AAA+ domain of Orc3. Although the Orc1/Orc4 active site is required for activity, in the D. melanogaster structure Orc1 is disengaged from Orc4 and positioned above the plane of the AAA+ ring (PDB = 4XGC). A color version of this figure is available online.
S. cerevisiae Cdc6 forms a stable complex with ORC in the presence of DNA and ATP (Wang et al., 1999, Mizushima et al., 2000, Seki and Diffley, 2000, Speck et al., 2005, Randell et al., 2006). By using a non-hydrolyzable ATP analog, a stable Cdc6•ORC complex can be trapped in the absence of DNA (Speck et al., 2005). Low-resolution 3D electron microscopy reconstructions of eukaryotic ORC have revealed an elongated, crescent-shaped particle that, in the presence of Cdc6 and ATPγS, transforms into a closed ring with a large central cavity (Speck et al., 2005, Clarey et al., 2006, Clarey et al., 2008, Sun et al., 2012). Structural investigation of D. melanogaster ORC has demonstrated a subunit order of Orc1→Orc4→Orc5→Orc3→Orc2 around the ORC ring (Bleichert et al., 2015), with a physical gap between the terminal Orc subunits that accommodates Cdc6 and is consistent with the direct interaction observed between Cdc6 and Orc1 (Sun et al., 2012, Sun et al., 2013, Wang et al., 1999). The C-terminus of Orc6 has been found to tether this subunit to a conserved domain insertion within Orc3 (Bleichert et al., 2013, Bleichert et al., 2015); S. cerevisiae Orc6 also binds Orc2 (Chen et al., 2008, Sun et al., 2012), possibly through a region that is specific to fungal homologs (Bleichert et al., 2013). The adjoining nature of Orc1 and Orc4 within the ternary complex is consistent with the known formation of a joint ATPase site between the two subunits (Klemm et al., 1997, Chesnokov et al., 2001, Bowers et al., 2004, Giordano-Coltart et al., 2005).
The ORC•Cdc6 assembly represents the functional initiator complex at origins. Interestingly, ORC itself exhibits differing levels of stability across species, suggesting that, in certain cases, ORC subcomplexes may be sequentially recruited. Indeed, unlike D. melanogaster and S. cerevisiae ORC, which form stable heterohexamers (Bell and Stillman, 1992, Chesnokov et al., 1999), human and X. laevis ORC appear to have alternative core subcomplexes. Human Orc1 and Orc6 loosely associate with an Orc2-5 core (Dhar et al., 2001, Vashee et al., 2001, Siddiqui and Stillman, 2007), whereas in X. laevis, the Orc6 subunit is labile (Gillespie et al., 2001). The weak interactions of certain Orc subunits likely play an important role in regulation. For example, vertebrate Orc1 is selectively released from chromatin after initiation (Rowles et al., 1999, Natale et al., 2000, Li et al., 2004), which helps to prevent re-initiation and provides a means to alter origin usage in a developmental- or differentiation-dependent fashion (Li and DePamphilis, 2002). The ability of either Orc1 or Orc6 to dissociate from ORC without disrupting the remaining core complex is consistent with the terminal position of Orc1 within the core ring and the peripheral binding site of Orc6; one exception to this trend occurs in S. pombe, whereby an ORC pentamer can be purified that lacks Orc4, an internal component of the ORC ring (Moon et al., 1999, Kong and DePamphilis, 2001). Whether S. pombe ORC retains additional stabilizing elements that can compensate for the absence of Orc4 in such instances is not known.
The recent atomic-resolution structure of D. melanogaster ORC (Bleichert et al., 2015) has helped clarify both our understanding of ORC organization and also mechanistic models for origin engagement and helicase loading. ORC adopts a two-tiered, notched ring architecture in which the WH domains of Orc1-5 sit atop a layer of AAA+ subunits (Figure 3B). Interestingly, the arrangement of AAA+ and WH domain contacts is domain swapped in ORC, such that the WH domain of Orc1 rests on the AAA+ region of Orc4, the WH domain of Orc4 sits on the AAA+ region of Orc5, and so forth. The open-ended, pentameric AAA+ core of ORC in principle provides an opportunity for the formation of four bipartite AAA+ interfaces. Within the D. melanogaster ORC structure, however, only one of the observed subunit interfaces evinces a typical AAA+ active-site configuration (Orc4/Orc5). Two others (Orc5/3 and Orc3/2) approximate the correct protomer alignment but fail to reconstitute a canonical active site; in the case of Orc3/2, the two subunits lack the amino acids necessary to bind ATP, as predicted. Of the remaining possible AAA+ interactions, Orc1 was unexpectedly found to be rotated more than 90 degrees out-of-plane from Orc4, rendering this catalytic center inoperative (Figure 3B). This finding was surprising, as Orc1 alone is capable of both binding and hydrolyzing ATP, using an arginine finger donated by Orc4 (Klemm et al., 1997, Chesnokov et al., 2001). Given that the D. melanogaster Orc1 conformation seen crystallographically is also seen in 3D electron microscopy reconstructions of ORC (Bleichert et al., 2013, Bleichert et al., 2015), this observation suggests that metazoan ORC can transition between at least two conformations, an autoinhibited and active conformation. In the future, it will be important to understand how the cell regulates the equilibria of ORC between these states, as well as whether this conformational transition is preserved in other ORC homologs.
A two-state model for ORC origin recognition
Despite the fundamental role of ORC in origin selection and recognition, the mechanism by which ORC associates with DNA has remained highly enigmatic. The presence of numerous interspecies peculiarities, such as divergent origin features, species-specific DNA-binding elements, and the effect of chromatin-bound trans-acting factors, have all challenged our understanding of how ORC is recruited to and stably binds origins. However, an analysis of ORC behavior across species suggests that in all cases, the origin-binding properties of ORC can be interpreted within a model containing at least two states: a transient ORC recruitment event that is mediated through diverse and sometimes species-specific interactions, and a second, mechanistically conserved step that positions ORC for productive origin engagement and that leads to stable Cdc6 association. We will first discuss the conserved mechanism by which ORC stably associates with origins and then detail the interactions that facilitate ORC recruitment.
Structural studies of origin-bound archaeal Orc have revealed a coordinated role for both the AAA+ and WH domains in DNA binding (Gaudier et al., 2007, Dueber et al., 2007). The WH domains show extensive contacts with origin DNA by a canonical helix-turn-helix (HTH) and β-hairpin wing interface, and a near compete loss of DNA binding by archaeal Orc is observed upon mutation or deletion of this region (Singleton et al., 2004, Dueber et al., 2011). The WH-DNA interaction positions the AAA+ domain in an orientation where a characteristic α-helical insertion within the initiator/helicase loader subgroup of AAA+ proteins, the initiator specific motif (ISM), contacts DNA. Given the different subunit compositions and oligomeric states observed between archaeal (monomer) and eukaryotic initiators (heterohexamer) it was initially unclear to what extent the mechanism of archaeal origin binding would be conserved. However, superposing the structure of DNA-bound archaeal Orc onto the D. melanogaster ORC crystal structure reveals nearly perfect co-axial positioning of the DNA within the ORC central channel (Bleichert et al., 2015). Notably, the eukaryotic ISMs and the wings of the WH domains each form a contiguous DNA-binding surface that lines the ORC central channel, and are thus positioned to engage DNA in a manner similar to that of archaeal Orc (although not identically, as the HTH motif of ORC’s WH domains are buried between inter-subunit contacts in the complex (Bleichert et al., 2015)). This modeling also accounts for an unresolved region of density in an early 3D electron microscopy reconstruction of an ORC•Cdc6•DNA complex (Sun et al., 2012). Overall, these data indicate that archaeal and eukaryotic ORC engage DNA by a generally conserved mechanism.
The insights gleaned from the available structural studies suggest a conserved mechanism for stable origin association by eukaryotic ORC. In this model, duplex DNA is loaded laterally into the ORC central channel through the discontinuity between Orc1 and Orc2 (Speck et al., 2005). This gap is then blocked off by the subsequent recruitment of Cdc6, generating a stable, closed-ring conformation that encircles origin DNA (Bleichert et al., 2015). Domain swapping between the WH domain of Orc2 and the AAA+ region of Cdc6 (and between the WH domain of Cdc6 and the AAA+ element of Orc1) are predicted to form, stabilizing the complex (Bleichert et al., 2015). The ability of the ORC•Cdc6 complex to encircle DNA would be predicted to underpin ORC’s observed persistence at origins (Speck et al., 2005, Duzdevich et al., 2015) and is consistent with the observation that both yeast and metazoan Cdc6 can stabilize ORC on DNA (Harvey and Newport, 2003, Houchens et al., 2008).
Interestingly, an additional DNA-binding element has been identified between the N-terminal BAH and AAA+ domain of S. cerevisiae Orc1 that is essential for ARS binding (Kawakami et al., 2015). While the related residues are largely conserved in metazoan Orc1, they are positioned outside of the ORC central pore and thus it is unclear how this interaction is coordinated with the encirclement mechanism described here, or whether it contributes only to the initial recruitment of ORC to DNA.
Diverse mechanisms for origin recruitment of ORC
Although the encirclement model accounts for stable origin binding, eukaryotic ORC has been observed to interact with DNA in a mode that exhibits fast kinetics (Harvey and Newport, 2003, Remus et al., 2004, Houchens et al., 2008, Duzdevich et al., 2015) and, in S. cerevisiae, allows for a one-dimensional linear search for bona fide origin sites (Duzdevich et al., 2015). This initial recruitment of ORC to DNA likely functions as an important intermediate on the path towards stable origin binding. Interestingly, substantial mechanistic plasticity appears to have been introduced to the recruitment step, co-evolving in certain cases with species-specific features.
In S. pombe ORC, the preference for AT-rich origins correlates with a unique domain insertion in Orc4 comprising nine AT-hook motifs (Chuang and Kelly, 1999, Moon et al., 1999), a DNA-binding element that facilitates interactions with the minor groove of AT-rich DNA sequences (Aravind and Landsman, 1998). Notably, the AT-hook motif is absent in S. cerevisiae and metazoan Orc4. Unlike S. cerevisiae ORC, which utilizes all AAA+ domain-containing Orc subunits for DNA binding (Lee and Bell, 1997), S. pombe Orc4 is necessary and sufficient for initial origin engagement (Chuang and Kelly, 1999, Moon et al., 1999, Kong and DePamphilis, 2001, Gaczynska et al., 2004), and the AT-hook motifs of Orc4 are additionally required for viability (Chuang et al., 2002). Interestingly, S. pombe ORC shows a biphasic mechanism of DNA binding, with an initial, salt-sensitive DNA binding event that precedes the formation of a salt-stable form, a state that in turn can be further stabilized by the addition of Cdc6 (Houchens et al., 2008). These findings suggest that S. pombe ORC is recruited to chromosomes by the Orc4 AT-hook motif, which then leads to stable DNA association through a mechanism that likely involves the encirclement of duplex DNA.
Analogous to S. pombe Orc4, metazoan Orc6 has been shown to have a distinct DNA binding activity. Despite the absence of an ATPase or WH domain, exclusion of Orc6 from D. melanogaster ORC results in the loss of ATP-dependent DNA binding, the same effect observed for Orc1 Walker A (ATP binding) and Walker B (ATP hydrolysis) mutants (Chesnokov et al., 2001). Analysis of the D. melanogaster and human Orc6 N-terminus reveals structural homology with the DNA-binding domain of transcription factor TFIIB (Chesnokov et al., 2003, Balasov et al., 2007), and mutation of the Orc6 TFIIB domain abolishes DNA binding (Liu et al., 2011). Although the Orc6 TFIIB domain was initially considered unique to metazoans, subsequent sequence analysis indicates that the domain is conserved in fungal Orc6, but lacks specific DNA-binding elements (Bleichert et al., 2013), a finding consistent with S. cerevisiae Orc6 being dispensable for origin recognition in vitro (Lee and Bell, 1997, Chen et al., 2007). Given the available data, it seems likely that metazoan ORC utilizes Orc6 to loosely tether the complex to DNA in a functionally analogous manner as S. pombe Orc4, and that this action aids with the initial ORC recruitment event that precedes stable origin association. This model begs the question, however, of how ORC is recruited to chromosomes in species where Orc6 is a labile subunit (Dhar et al., 2001, Vashee et al., 2001, Gillespie et al., 2001). One possible answer is that chromatin-bound Orc6 may function as a recruitment platform for the core Orc1-5 subunits.
In addition to “hard-wired” DNA binding domains, eukaryotic ORC can also bind a plethora of chromatin-associated factors that provide additional means for recruiting and regulating ORC’s association with origins (Chakraborty et al., 2011). Early studies recognized that ORC played dual roles in replication and transcriptional regulation (Bell et al., 1993, Loo et al., 1995, Palacios DeBeer et al., 2003, Leatherwood and Vas, 2003), and a number of transcriptional regulators have been shown to bind ORC, including HP1, E2F, HMGA1a and Sir1 (Triolo and Sternglanz, 1996, Pak et al., 1997, Fox et al., 1997, Gardner et al., 1999, Royzman et al., 1999, Bosco et al., 2001, Prasanth et al., 2004, Thomae et al., 2008, Prasanth et al., 2010). Although chromatin accessibility and transcriptional programs are clearly regulated in an ORC-dependent manner (Bell et al., 1993, Foss et al., 1993, Micklem et al., 1993, Loo et al., 1995, Fox et al., 1995, Pak et al., 1997, Huang et al., 1998, Bose et al., 2004, Chesnokov, 2007, Shor et al., 2009), whether chromatin bound transcriptional regulators also direct ORC origin usage is less clear. Certain cases have been investigated in some detail; for example, ORC can be targeted to chromatin-bound HMGA1a, where it directs assembly of replication complexes (Thomae et al., 2008), while genome-wide analysis of metazoan origins reveals ORC association with HP1 sites (Cayrou et al., 2011). Thus, transcriptional regulators represent an additional means of tethering ORC to chromosomes and likely affect ORC function in replication.
Eukaryotic ORC is also directly recruited to histones by the N-terminal Bromo-Adjacent Homology (BAH) domain in Orc1 (Zhang et al., 2002, Noguchi et al., 2006, Muller et al., 2010). This element selectively binds H4K20me2, an interaction that directs replication licensing (Kuo et al., 2012). Other histone modifications also have been found to direct origin usage (such as H4K20me1, H3K4me2/3, and H4 acetylation), but how these effect ORC positioning or downstream licensing reactions remains unclear (Rice et al., 2002, Vogelauer et al., 2002, Aggarwal and Calvi, 2004, Knott et al., 2009, Tardat et al., 2010, Miotto and Struhl, 2010, Costas et al., 2011, Eaton et al., 2011, Liu et al., 2012b). A factor known as ‘ORCA’ (for Origin Recognition Complex Associated) additionally has been reported to directly recruit human ORC to chromosomes, as well as to bind to other initiation factors that promote replication licensing and S-phase progression (Shen et al., 2010, Shen et al., 2012). Conversely, the assembly of the replication machinery at telomeres has been reported to be controlled by an interaction between ORC and the TRF2 subunit of the shelterin complex (Atanasiu et al., 2006, Tatsumi et al., 2008, Deng et al., 2009, Higa et al., 2016).
Collectively, multiple lines of data demonstrate that localizing ORC to DNA prior to productive origin engagement (i.e., Cdc6-dependent DNA encircling) is a critical step in replication initiation, and that many diverse recruitment mechanisms are sufficient to demarcate a replication start site. Consistent with this idea is the demonstration that a fusion between ORC and the Gal4 DNA binding domain can result in the initiation of DNA replication on a plasmid containing a tandem array of Gal4 binding sites (Takeda et al., 2005). In light of the many mechanisms that can facilitate ORC recruitment, an important future direction will be to understand how different recruitment pathways cooperate or antagonize each other in specifying sites of replisome assembly.
The Minichromosome Maintenance (MCM) Complex
Following the active designation of an origin by ORC, the complex next facilitates initiation at these sites by loading the heterohexameric MCM2-7 helicase onto DNA. Studies into MCM2-7 function have revealed a surprisingly complex enzyme that appears capable of harnessing ATP to promote both the melting of double-stranded DNA and translocation of single-stranded DNA substrates. Although precise mechanisms have yet to be elaborated, a picture of how specific MCM2-7 elements couple ATP turnover to DNA remodeling and movement is beginning to emerge.
Identification of the MCM2-7 complex
MCM proteins were originally identified in genetic screens that aimed to uncover factors required for replication and cell cycle progression in yeast (Maine et al., 1984, Yan et al., 1991, Chen et al., 1992, Hennessy et al., 1990, Moir et al., 1982). Homologs were subsequently identified in D. melanogaster (Treisman et al., 1995), X. laevis (Madine et al., 1995) and in humans (Todorov et al., 1995), and like yeast MCMs, were shown to have an essential role in DNA replication. Investigation into MCM function converged with studies by Laskey and Blow, who identified a replication licensing factor (RLF) that restricted replication to once per cell cycle by nuclear exclusion until after nuclear envelope breakdown (Blow and Laskey, 1988, Blow, 1993). Indeed, MCM proteins, like the RLF, showed redistribution from the cytosol to nucleus upon completion of mitosis (Hennessy et al., 1990), and immunodepletion of MCM proteins inhibited replication (Kubota et al., 1995, Madine et al., 1995). Importantly, a purified RLF fraction was found to resolve into two separate factors, RLF-M and RLF-B, with RLF-M containing multiple proteins that cross-reacted with MCM antibodies (Chong et al., 1995).
Purification of RLF-M, as well as co-immunoprecipitation studies, revealed that MCM proteins reside within large multimeric assemblies with other MCM2-7 family members (Kubota et al., 1995, Madine et al., 1995, Chong et al., 1995, Burkhart et al., 1995, Kimura et al., 1995, Musahl et al., 1995, Romanowski et al., 1996a, Chong et al., 1996). The predominant assembly in vivo is an MCM2-7 heterohexamer, although low levels of subassemblies have been reported to exist that may represent intermediates (Su et al., 1996, Ishimi, 1997, Adachi et al., 1997, Thommes et al., 1997, Lee and Hurwitz, 2000, Prokhorova and Blow, 2000). As with Orc1-5, MCM proteins are predicated upon a conserved AAA+ ATPase element (Koonin, 1993, Iyer et al., 2004). However, unlike Orc1-5, which show differing levels of active site conservation, the six subunits in the MCM2-7 complex each contain the complete set of catalytic residues expected to be necessary for supporting ATP hydrolysis.
MCM architecture
The basic architecture of an MCM protein is conserved across archaea and eukaryotes. MCMs contain a central AAA+ ATPase fold flanked by conserved N- and C-terminal domains (termed ‘NTD’ and ‘CTD,’ respectively). Structural analyses of MCMs have revealed three conserved subdomains within the NTD, allowing subdivision of this element into the NTD-A, -B, and -C regions (Figure 4A) (Fletcher et al., 2003, Liu et al., 2008, Brewster et al., 2008, Bae et al., 2009, Fu et al., 2014, Li et al., 2015). NTD-C, the most highly conserved NTD subdomain, possesses an oligonucleotide/oligosaccharide-binding (OB) fold. Within an MCM hexamer, interactions between adjacent NTD-C elements serve as the primary interface for subunit assembly, with this interaction alone sufficient to facilitate hexamerization of the NTD (Fletcher et al., 2003, Kasiviswanathan et al., 2004). Although NTD-A is the least conserved N-terminal subdomain, it forms a helical bundle with similarity to helix-turn-helix type DNA binding proteins (Costa et al., 2008). The NTD-B element comprises a Zn-finger motif that, with the exception of MCM3, is universally conserved in MCM2-7 (Li et al., 2015), and that facilitates higher order organization of single NTD hexamers into a conserved, head-to-head double hexamer (Figure 4B) (Kelman et al., 1999, Shechter et al., 2000, Chong et al., 2000, Fletcher et al., 2005, Remus et al., 2009, Evrin et al., 2009, Costa et al., 2014, Li et al., 2015).
Figure 4.
MCM architecture. A) MCM homologs are characterized by three domains: NTD, AAA+, and CTD. The NTD can be subdivided into NTD-A (a small helical bundle), NTD-B (Zn-finger), and NTD-C (OB fold). The CTD forms a WH domain (for AAA+ and NTD, PDB = 3F9V; for WH domain, PDB = 2KLQ). B) Two physiologically relevant MCM oligomers have been observed, a hexamer that is formed by lateral interactions between the AAA+ and NTD domains of adjacent protomers, and a double hexamer that is formed by interactions between the NTD-B Zn-finger domains of two MCM2-7 rings. The double hexamer structure from S. cerevisiae Mcm2-7 is shown (Li et al., 2015), with one hexamer faded compared to the other. The inset shows a top-down view through the central cavity and the radial arrangement of eukaryotic MCM2-7 subunits (the double hexamer was built from two copies of PDB = 3JA8 fit to the EM density map EMD-6338 (Li et al., 2015)). A color version of this figure is available online.
The AAA+ domain is the most highly conserved region across MCM homologs. The MCM AAA+ fold contains three unique insertions relative to prototypical AAA+ proteins, and falls within the pre-sensor II (PSII) insert clade of the ATPase superfamily (Iyer et al., 2004, Erzberger and Berger, 2006). One such insertion is distinctive in that it remodels a portion of the AAA+ cassette to orient a catalytic amino acid known as the sensor II motif in trans with the active site of an adjacent protomer (Moreau et al., 2007, Bae et al., 2009), rather than in cis, which serves as the more usual arrangement in AAA+ oligomers. By comparison, the other two insertions are found in many other AAA+ subgroups and consist of two β–hairpin insertions termed the pre-sensor I (PSI)-insert and the helix 2 (H2)-insert. Within the context of the hexamer, these two β-hairpins are positioned within the central channel and are integral to helicase mechanism (Figure 5). With regard to the MCM catalytic centers, a recent high-resolution cryo-electron microscopy structure of a full-length S. cerevisiae MCM2-7 double hexamer reveals significant structural variability between the six radially-arranged ATPase sites, with a catalytically competent conformation but unequal nucleotide occupancy observed for four of the sites (Li et al., 2015). This variability is consistent with the unequal ATPase activity observed for the six MCM2-7 active sites (Bochman et al., 2008) and with the non-equivalent function of these sites across the different stages of helicase recruitment, loading, and activation (Ilves et al., 2010, Coster et al., 2014, Kang et al., 2014).
Figure 5.
Functional elements of MCM helicases. A) Each MCM monomer contains multiple functional elements, including DNA-binding/sensing motifs, regions that modulate ATPase activity, and loops that communicate between the NTD and AAA+ domain (PDB = 3JA8, chain 2). B) In the context of a hexamer, the MCM functional elements (excepting the external β-hairpin) line a central cavity through which DNA translocates (modeled after PDB = 3JA8, chains 4, 6, and 7). C) Symbol key. D) Detailed functional description for each MCM element known to contribute to activity. A color version of this figure is available online.
Insofar as the MCM CTD, solution structures have shown this region to adopt a canonical winged-helix (WH) domain (Figure 4A) (Wei et al., 2010, Liu et al., 2012a, Wiedemann et al., 2015), which based on the weak or absent CTD density in structures of many full-length MCM proteins, appears flexibly tethered to the AAA+ core. Although the position of this domain with respect to the hexamer was at first unclear, recent structural and biochemical work demonstrates that the WH domain sits distal to the NTD (Li et al., 2015, Wiedemann et al., 2015, Yuan et al., 2016). The S. solfataricus MCM WH domain can bind single-stranded DNA weakly (Pucci et al., 2007); however, what role this activity plays in vivo, and whether the WH domains of the eukaryotic MCMs can also bind DNA, are open questions. The majority of data reported thus far suggests that, at least in archaea, the MCM WH domain functions to allosterically modulate the ATP hydrolysis rate of the AAA+ domain (Jenkinson and Chong, 2006, Barry et al., 2007, Wiedemann et al., 2015). In addition, this region can serve as a protein/protein interaction site, with the WH domains of eukaryotic MCM3 and MCM6 having been shown to engage Cdc6 and Cdt1, respectively (You and Masai, 2008, Liu et al., 2012a, Frigola et al., 2013). Interestingly, recent structural analyses of the activated yeast helicase (the CMG complex) have revealed that the Mcm5 and Mcm6 WH domains partially occlude the Mcm2-7 central channel, thus positioning them to perhaps function in the translocation process (Yuan et al., 2016). Despite this observation, the precise role of the MCM WH domains remains to be determined.
Although the ATPase activity of MCM proteins localizes exclusively to the core AAA+ domain, hexamer formation is facilitated by both the AAA+ and NTD elements in the order MCM5→MCM3→MCM7→MCM4→MCM6→MCM2 (Schwacha and Bell, 2001, Crevel et al., 2001, Davey et al., 2003, Costa et al., 2011). A loop within each NTD-A OB-fold, termed the allosteric communication loop (ACL), contacts the AAA+ domain of its adjacent protomer, further stabilizing the hexamer (Miller et al., 2014, Li et al., 2015). Collectively, these interactions facilitate the formation of a particle consisting of two stacked rings (a AAA+ tier and an NTD tier). In single and double MCM2-7 hexamers, the central channel is sufficiently large enough to accommodate double-stranded DNA (Figure 4B) (Remus et al., 2009, Evrin et al., 2009, Costa et al., 2011, Lyubimov et al., 2012, Costa et al., 2014, Li et al., 2015); later, in an unknown series of events, the helicase undergoes an isomerization reaction that appears coupled to the extrusion of one of the two strands through a natural discontinuity between the MCM2/5 subunits (Bochman and Schwacha, 2008, Costa et al., 2011, Bruck and Kaplan, 2015b), permitting translocation along single-stranded DNA (Ishimi, 1997, Kelman et al., 1999, Shechter et al., 2000, Ilves et al., 2010, Fu et al., 2011).
MCM mechanism in origin melting and processive unwinding
The MCM2-7 complex matures through a variety of intermediates before being incorporated into the active helicase present at replication forks. Within the MCM2-7 lifecycle, two stable complexes are observed: an origin-bound double hexamer (Evrin et al., 2009, Remus et al., 2009) and a replication fork-associated single hexamer (Fu et al., 2011, Duzdevich et al., 2015). Interestingly, the double hexamer serves not only as a platform for recruiting helicase-activating factors but also is likely responsible for the initial origin melting event (Sun et al., 2014, Li et al., 2015, Bochman and Schwacha, 2015). Thus, the MCM complex must utilize the ATP-hydrolysis-driven repositioning of its DNA-binding elements to carry out two very distinct tasks, origin opening as a double hexamer and processive DNA unwinding as a single hexamer. Although the MCM2-7 DNA-binding elements are relatively well defined, it is currently unclear whether the two functionalities of the helicase require an overlapping or mutually exclusive set of protein interactions with DNA.
The MCM2-7 hexamer has multiple tiers of DNA-binding elements that are positioned to engage DNA passing through the central pore (Figure 5). The NTD-B Zn-fingers are situated at one end of the hexamer, forming a skirt of DNA-binding elements that run parallel to and extend the central channel. This domain is required for double hexamer formation, and mutation or deletion of the archaeal NTD-B reduces DNA binding and results in loss of helicase activity (Poplawski et al., 2001, Pucci et al., 2004, Kasiviswanathan et al., 2004). The Zn-finger motifs of eukaryotic MCM2-7 also facilitate double-hexamer formation (Li et al., 2015) and represent a putative DNA-binding region, but the role of this domain in helicase activity has not been investigated directly. The Zn-finger motif of S. cerevisiae Mcm2 and Mcm5 are required for cellular proliferation, suggesting an essential and conserved function for this region (Yan et al., 1991, Dalton and Hopwood, 1997).
As one moves through the central MCM2-7 pore, from the NTD to the CTD, the next DNA-binding elements encountered are the β–hairpin insertions within the OB-fold (NTD-C). The NTD-C β–hairpin constricts the central channel (Fletcher et al., 2003, Li et al., 2015) and, in archaea, is electropositive in nature. Mutation of basic residues within this loop and on the adjacent, pore-lining surface of the OB-fold dramatically reduce both DNA binding and helicase activity in archaeal MCM (Fletcher et al., 2003, Pucci et al., 2004, McGeoch et al., 2005, Fletcher et al., 2008). The NTD-C β-hairpin is highly variable in length and composition within the eukaryotic MCM2-7 family, and while a majority are enriched in basic residues, this conservation is not universal. Nevertheless, the NTD-C region appears critical for eukaryotic MCM function, as S. cerevisiae strains harboring mutations within this β–hairpin of Mcm5 show an increased rate of minichromosome loss (Leon et al., 2008). This effect can be repressed by the addition of multiple ARS sequences, which suggests that origin recruitment, but not helicase activity, is defective in this particular mutant.
In addition to the NTD-C β–hairpin, the crystal structure of an archaeal MCM NTD bound to single-stranded DNA has revealed a second DNA-binding region within the NTD-C, the MCM-single-stranded-DNA binding motif (MSSB) (Froelich et al., 2014). The MSSB is formed primarily by two positively charged residues that extend from the OB-fold and, interestingly, bind single-stranded DNA in an orientation perpendicular to the long axis of the central channel. The MSSB DNA-binding residues are conserved among three consecutive subunits of the eukaryotic MCM2-7 complex (MCM4, MCM6, and MCM7), but not the other subunits. It has been suggested that the MSSB plays a role in DNA melting, working against a DNA pumping action of the ATPase domains to induce topological strain that encourages strand separation (Froelich et al., 2014). Thus, the MSSB may have a role in the early steps of initiation, consistent with the defects observed for S. cerevisiae MSSB mutants in in vitro loading reactions (Froelich et al., 2014); whether or how this motif contributes to helicase function during elongation is unknown.
The AAA+ PSI β–hairpin insertion constitutes the most C-terminal DNA-binding element. This pore-lining loop shows the highest level of conservation of the various DNA binding elements discussed thus far, and contains an invariant lysine that projects towards the central pore. The PSI β-hairpin has been compared to the translocation β–hairpin of SF3 helicases (e.g., the papillomavirus E1 protein and the SV40 Large T-antigen), which in the context of a hexamer forms a vertically aligned, right-handed staircase that tracks the DNA backbone with a conserved lysine (Enemark and Joshua-Tor, 2006, Enemark and Joshua-Tor, 2008). Consistent with a critical role in helicase function, mutation of the PSI lysine in archaeal MCM weakens DNA binding and ATP hydrolysis, and fully abrogates helicase activity (McGeoch et al., 2005); mutation of this residue within the context of the D. melanogaster CMG similarly abolishes helicase function (Petojevic et al., 2015). Interestingly, genetic analysis of the S. cerevisiae PSI β–hairpin lysines has highlighted non-equivalent functions for this residue in the six Mcm2-7 homologs, with only the mutation of this residue in Mcm3 abolishing an ability to complement deletion strains (Lam et al., 2013, Ramey and Sclafani, 2014).
How are the activities of the MCM2-7 DNA-binding elements functionally coordinated with ATP turnover to achieve the different functionalities observed for the helicase? At least two other loops seem to facilitate communication between the AAA+ domain and the DNA-binding elements lining the central channel. One is a motif known as the H2-insert, which sits in the MCM AAA+ fold and forms an extended loop that junctions the ATPase and NTD tiers, creating a pore-lining feature at the interface (Brewster et al., 2008, Bae et al., 2009, Li et al., 2015). Notably, analysis of the current data suggests that ATP-dependent repositioning of the H2-insert may switch the helicase between an origin melting conformation to one that facilitates processive unwinding. Despite a low level of sequence conservation, the length of the H2-insert is fully conserved in archaea and all MCM2-7 homologs and is sufficiently long to directly contact the MSSB. Interestingly, the H2-insert is rich in charged amino acids and alignment of the six yeast Mcm proteins reveals two fully conserved acidic residues that, in the context of the double hexamer, are positioned to shield the MSSB from binding DNA (Figure 5). Consistent with this proposal, the H2-insert dramatically affects DNA binding in the archaeal MCM complex, with removal of the loop enhancing MCM affinity for both double and single-stranded DNA (Jenkinson and Chong, 2006). The position of the H2-insert is thought to be modulated in an ATP-dependent fashion, such that the presence of ATP results in a more buried state for this loop (Jenkinson and Chong, 2006). In conclusion, the available data suggest a critical role for the H2-insert in transitioning the helicase between different functional states.
In addition to H2-insert dependent crosstalk between MCM subunits, the ACL of NTD-C projects upward from the OB-fold and toward the AAA+ domain of an adjacent protomer, where it is sandwiched between the H2-insert and PSI β–hairpin (Sakakibara et al., 2008, Barry et al., 2009, Li et al., 2015). Like the other functional motifs within the NTD, the ACL is absolutely required for helicase activity. However, the ACL does not affect DNA binding activity, but instead modulates the ATPase activity of the AAA+ domain; mutation of ACL residues or deletion of the region results in markedly reduced ATPase activity and inhibition of duplex unwinding (Sakakibara et al., 2008, Barry et al., 2009). Like the H2-insert, the ACL position with respect to the AAA+ domain is altered under different nucleotide-bound states (Barry et al., 2009).
Overall, the DNA-binding elements within the MCM NTD appear to play a critical function in both origin melting and unwinding. Although functional data point to an ATP-controlled connection between the positional status of specific DNA binding elements within the MCM2-7 ring and their ability to grasp or release substrate, direct observation of these elements in either a DNA melting or translocation mode is currently lacking. Recent electron microscopy models of the activated eukaryotic MCM2-7 helicase in the context of the CMG have revealed that there exists conformational coupling between the NTD and CTD tier, suggestive of coordinated action between these regions during processive unwinding (Yuan et al., 2016, Abid Ali et al., 2016). NTD-A also exhibits conformational dynamics and has been seen to rotate markedly away from the hexamer axis, a movement that may be coordinated through direct binding of DNA to the NTD-A (Fletcher et al., 2003, Chen et al., 2005, Hoang et al., 2007, Liu et al., 2008, Costa et al., 2008). As additional substrate-bound structures are solved for the archaeal and eukaryotic MCM homologs, exciting insights into its DNA remodeling and motor mechanisms are certain to emerge.
Loading of the replicative helicase
The defining step in eukaryotic replication licensing is loading of MCM2-7 onto duplex DNA into a stable head-to-head double hexamer (Figure 1) (Remus et al., 2009, Evrin et al., 2009, Gambus et al., 2011). To attain this state, single MCM2-7 hexamers are sequentially loaded by an interaction with ORC•Cdc6 that is chaperoned by the Cdc10-dependent transcript 1 (Cdt1) protein. Structural characterization of multiple intermediates, together with the ability to reconstitute and study loading in vitro, are revealing how the stepwise and carefully orchestrated exchange of pre-RC factors at origins underlies this complex reaction (Figure 6).
Figure 6.
MCM2-7 complex loading and maturation into the CMG. Two sequential rounds of helicase recruitment and loading at origins is required for building an MCM2-7 double-hexamer. For both hexamers, DNA is threaded into the central channel through a discontinuity between MCM2 and MCM5. The first hexamer is recruited through direct interactions with the initiator (MCM3-Cdc6) and may require Cdt1 for overcoming an MCM6-mediated autoinhibited state of the helicase. After the first hexamer loads, both Cdc6 and Cdt1 are released. Rebinding of Cdc6 to ORC primes the system for recruiting and loading a second hexamer in the opposite direction as the first, an event that has been proposed to be controlled by ORC•Cdc6, but templated by the first MCM2-7 hexamer. Cdt1 and Cdc6 recruitment and ejection are required for both loading events. Phosphorylation of the double hexamer and other initiation factors by CDK and DDK facilitate origin melting, GINS and Cdc45 recruitment/assembly, DNA strand extrusion, and activation of the helicase for DNA unwinding. A color version of this figure is available online.
Mechanism of helicase recruitment to origins
Maturation of the pre-RC occurs in a sequential fashion during the G1 phase of the cell cycle (Romanowski et al., 1996b, Gillespie et al., 2001, Tsuyama et al., 2005, Remus et al., 2009, Evrin et al., 2009, Tsakraklides and Bell, 2010, Ticau et al., 2015). Demarcation of replication start sites by ORC represents the inaugurating event, and permits the recruitment of and stable association with Cdc6 (Liang et al., 1995, Cocker et al., 1996, Coleman et al., 1996, Donovan et al., 1997, Seki and Diffley, 2000, Tsuyama et al., 2005). In an in vitro setting, Cdc6 regulates the fidelity of origin selection by two distinct mechanisms. First, Cdc6 helps to restrict the initiator from acting at illegitimate origins by sequestering the free initiator (thereby lowering the effective concentration of the initiator to increase origin specificity) (Duzdevich et al., 2015), and by triggering the dissociation of ORC from non-origin DNA (Mizushima et al., 2000). Second, Cdc6 ATPase activity is enhanced when bound to non-ARS sequences, which triggers the dissociation of Cdc6 from the initiator (Speck et al., 2005, Speck and Stillman, 2007). Ultimately, the secure association between Cdc6 and ORC at origins primes the initiator for helicase recruitment and loading (Rowles et al., 1996, Donovan et al., 1997, Perkins and Diffley, 1998, Feng et al., 2000, Tsuyama et al., 2005); this same complex will later facilitate the release of the double hexamer for subsequent activation at the onset of S phase (Chang et al., 2015).
The MCM2-7 helicase must first be recruited to origins to initiate the loading reaction. Notably, the helicase is sensitive to solution conditions and shows substantial conformational heterogeneity (Gomez-Llorente et al., 2005, Bochman and Schwacha, 2008, Costa et al., 2011), a feature that is pertinent for understanding the mechanism of association with the origin-bound initiator. Biochemical and structural analyses of the MCM2-7 complex has revealed that, in the absence of ATP, archaeal, fungal, and metazoan complexes can adopt a cracked ring architecture that bears a discontinuity, or ‘gate,’ between MCM2 and MCM5 (Bochman and Schwacha, 2008, Costa et al., 2011, Lyubimov et al., 2012, Boskovic et al., 2016, Samson et al., 2016). Nucleotide binding constricts the ring and favors the formation of MCM2/5 interactions across the ring break (Samel et al., 2014), but does not appear to irreversibly close the gate in single MCM2-7 hexamers (Costa et al., 2011, Lyubimov et al., 2012). Thus, MCM2-7 may not require active ring-opening prior to loading, but instead may simply be stably aligned around DNA by origin-bound ORC•Cdc6.
In addition to the ORC•Cdc6 helicase-loader complex, Cdt1 is also required for helicase loading in yeast and metazoa (Maiorano et al., 2000, Nishitani et al., 2000, Whittaker et al., 2000, Claycomb et al., 2002, Devault et al., 2002). However, the role of this factor has been enigmatic. Although Cdt1 was initially thought to be required for helicase recruitment to ORC (Asano et al., 2007, Chen et al., 2007, Chen and Bell, 2011), recent data suggest that MCM2-7 is recruited by direct interactions with the initiator (Fernandez-Cid et al., 2013, Frigola et al., 2013). With the exception of Orc6, all initiation factors discussed thus far possess a central AAA+ ATPase unit; Cdt1 represents an additional non-AAA+ protein involved in the process. Cdt1 was first identified in S. pombe, and through an ability to induce re-replication in the absence of cell division, was recognized as an essential and potent initiator of replication (Hofmann and Beach, 1994, Nishitani et al., 2000). Cdt1 is universally conserved in metazoa but is inconsistently found in other eukaryotic supergroups (Tada et al., 1999, Maiorano et al., 2000, Whittaker et al., 2000, Aves et al., 2012). The weak conservation of Cdt1 across species aggravated initial attempts to identify an S. cerevisiae homolog, but eventually the TAH11 (Topoisomerase-A hypersensitive 11) protein was identified as the budding yeast counterpart (Tanaka and Diffley, 2002, Devault et al., 2002). TAH11 was first isolated in genetic screens for genes that showed synthetic lethality with a mutant topoisomerase I allele (Fiorani and Bjornsti, 2000), and proteomic studies in human cells have suggested there exists an interaction between Cdt1 and both topoisomerase I and topoisomerase IIa (Sugimoto et al., 2008). Archaea also contain a Cdt1-related protein, termed WhiP; however, this factor appears to uniquely function in the Orc-independent assembly of pre-RCs (Robinson and Bell, 2007, Samson et al., 2013).
Interestingly, Cdt1 has markedly diverged between S. cerevisiae and metazoa, and functional analyses have suggested that there exist differing mechanisms for Cdt1-dependent action across species. All Cdt1 homologs are built from a conserved C-terminal pair of WH domains (Lee et al., 2004, Khayrutdinov et al., 2009). By contrast, the N-terminal region of Cdt1 is not conserved, and while the N-terminal domain of metazoa and S. pombe Cdt1 has a basic pI (human Cdt1 pI = 10.5) that is predicted to be unstructured, the S. cerevisiae N-terminus has an acidic pI (pI = 5.1) and possesses sufficient secondary structure to likely constitute an ordered domain (modeling using Phyre2 (Kelley et al., 2015) suggests that this region corresponds to a catalytically defunct oxygenase fold). Interestingly, an S. cerevisiae Cdt1 construct lacking the non-conserved N-terminal domain fails to load the helicase into an activation-competent conformation (Takara and Bell, 2011, Fernandez-Cid et al., 2013), whereas the N-terminus of metazoa Cdt1 is dispensable for the loading reaction (Ferenbach et al., 2005).
The factors with which Cdt1 stably associates also evince species-specific differences. In S. cerevisiae, Cdt1 and Mcm2-7 form a tight complex and exhibit interdependent nuclear import (Tanaka and Diffley, 2002, Kawasaki et al., 2006, Remus et al., 2009, Takara and Bell, 2011); Cdt1 similarly helps maintain the structural integrity of the Mcm2-7 complex (Wu et al., 2012, Frigola et al., 2013). Conversely, Xenopus Cdt1 and MCM2-7 are biochemically separable initiation factors, with immunodepletion of MCM3 from Xenopus egg extracts removing MCM2-7 but having no effect on Cdt1 (Maiorano et al., 2004) (it is worth noting that mouse Cdt1 has been reported to form a complex with MCM2-7 in vitro (You and Masai, 2008), but the stability of this assembly has not been investigated). Multiple explanations could account for the differences in behavior between S. cerevisiae and Xenopus Cdt1; for example, differences in the relative affinities of Cdt1 for the MCM2-7 complex could reflect an alteration to the sequential assembly of the pre-RC factors in metazoa, such that Cdt1 associates first with ORC•Cdc6 rather than with the helicase (Maiorano et al., 2000, Waga and Zembutsu, 2006). Finally, metazoa contain an additional Cdt1-binding protein, Geminin, which negatively regulates licensing by inhibiting the Cdt1-dependent loading of the helicase into a stable double hexamer (Wohlschlegel et al., 2000, Edwards et al., 2002, Yanagi et al., 2002, Wu et al., 2014a).
Despite the differences in Cdt1 structure and its preferred binding partners between budding yeast and metazoans, multiple studies have identified MCM6 as the primary Cdt1 interaction site on the helicase hexamer (Yanagi et al., 2002, Ferenbach et al., 2005, Zhang et al., 2010, Wei et al., 2010, Liu et al., 2012a, Wu et al., 2012) (mouse Cdt1 also interacts with MCM2 (You and Masai, 2008)). Cdt1/MCM6 contacts are facilitated by the C-terminal WH domain of each protein (Ferenbach et al., 2005, Teer and Dutta, 2008, Khayrutdinov et al., 2009, Jee et al., 2010, Wei et al., 2010, Takara and Bell, 2011), with structural studies of the human proteins revealing that a helical extension which projects from the Cdt1 WH region engages the helix-turn-helix motif of the MCM6 C-terminus (You and Masai, 2008, Liu et al., 2012a). Mutagenesis of residues in the interaction surface abolishes helicase loading and DNA replication in budding yeast, suggesting that the MCM6-Cdt1 mechanism of binding is widely conserved (Liu et al., 2012a). Despite this contact, in vitro binding studies have suggested that there exist additional undefined sites of interaction between Cdt1 and MCM2-7 (Khayrutdinov et al., 2009, Fernandez-Cid et al., 2013).
S. cerevisiae Cdt1 has also been reported to bind Orc6 (Semple et al., 2006, Asano et al., 2007, Chen et al., 2007, Chen and Bell, 2011). This interaction was initially proposed to facilitate helicase recruitment by helping to tether MCM2-7 to the ORC•Cdc6 complex (it is unclear whether metazoan Cdt1 and Orc6 interact (Yanagi et al., 2002)); however, other work has indicated that S. cerevisiae Mcm2-7 contains elements that directly engage ORC•Cdc6 and that are wholly sufficient for recruitment (Fernandez-Cid et al., 2013, Frigola et al., 2013). Cdt1’s participation in modulating MCM/ORC•Cdc6 interactions is still under debate, with one study noting that budding yeast Cdt1 relieves an Mcm6-dependent autoinhibitory mechanism that prevents Cdt1-independent recruitment (i.e., deletion of an Mcm6 autoinhibitory domain results in helicase recruitment in the absence of Cdt1) (Fernandez-Cid et al., 2013), and another showing that Cdt1 is fully dispensable for helicase recruitment (here only an interaction between Cdc6 and the extreme C-terminus of Mcm3 appears necessary for recruitment (Frigola et al., 2013). Consistent with the proposed role of the eukaryotic MCM3 C-terminus, recent studies in archaea have found that the MCM C-terminal WH domain directly binds Orc to facilitate recruitment of the helicase to replication origins (Samson et al., 2016).
Together, the available data suggest that Cdt1 may be dispensable for helicase recruitment, and that this step instead predominantly relies on interactions between ORC•Cdc6-bound origins and MCM2-7. What then might be the significance of Cdt1 and its interaction with Orc6? Beyond a potential ability to relieve MCM6-dependent autoinhibition (Fernandez-Cid et al., 2013), S. cerevisiae Cdt1 stabilizes Mcm2-7 at origins in an Orc6-dependent manner (Chen et al., 2007), such that in the absence of Cdt1, topologically linked Mcm2-7 hexamers fail to load (Chen and Bell, 2011). Notably, an N-terminal deletion construct of budding yeast Cdt1 permits loading of double hexamers, but in a state that is incapable of being subsequently activated (Takara and Bell, 2011). Thus, Cdt1 is required for events downstream of recruitment, which appear to include the formation of an MCM2-7 double hexamer intermediate that is competent for replication. Overall, the molecular details underlying the role of Cdt1, and in particular its Orc6 association, are in need of further study.
Loading the double hexamer
Ultimately, the initiator-dependent recruitment of MCM2-7 to origins leads to the formation of a stable chromatin-bound MCM2-7 complex that, unlike ORC•Cdc6, is resistant to high-salt extraction (Donovan et al., 1997, Edwards et al., 2002, Bowers et al., 2004). Notably, the MCM2-7 loading reaction has been fully reconstituted in vitro with purified proteins from S. cerevisiae (Remus et al., 2009, Evrin et al., 2009), which has in turn allowed the use of single molecule and structural techniques to understand both the nature of the salt-stable complex and the events leading to its formation. The details of the loading reaction discussed below are derived exclusively from studies with S. cerevisiae initiation factors; given some of the species-specific differences observed to date, it will be important to reconstitute the loading reaction in vitro with other organisms for comparison.
Following the recruitment of Cdt1•Mcm2-7 to ORC•Cdc6-bound origins, Mcm2-7 is deposited onto DNA while both Cdc6 and Cdt1 are ejected from the pre-RC in an ATP hydrolysis-dependent fashion. Cdc6 release occurs prior to Cdt1 both for single- and double-hexamer loading (Remus et al., 2009, Evrin et al., 2009, Tsakraklides and Bell, 2010, Kang et al., 2014, Ticau et al., 2015) (Figure 6). Singly-loaded Mcm2-7 hexamers result in the transient formation of a meta-stable ORC•Cdc6•Cdt1•Mcm2-7 (OCCM) intermediate, which can be stabilized by the presence of a non-hydrolyzable ATP analog. This property has been exploited to permit imaging of the complex by 3D cryo-electron microscopy (Sun et al., 2013). Within the OCCM, Cdt1 interfaces with the N-terminal tier of Mcm2, Mcm6, and Mcm5. The Mcm3 C-terminus resides proximal to Cdc6, consistent with the observed interaction between these regions in vitro and the proposed role of this interaction in helicase recruitment (Sun et al., 2013, Frigola et al., 2013). Cdt1 and Orc6 make no visible contacts in the OCCM structure; indeed the N-terminal TFIIB domain of Orc6 does not appear ordered. Notably, the Cdt1•Mcm2-7 heptamer is oriented such that the MCM C-terminal domains (the region containing the AAA+ and WH domains) abut the ORC WH domains (Bleichert et al., 2015), leaving the MCM NTD accessible. Although analysis of the loading reaction in bulk has indicated that singly-loaded Mcm2-7 hexamers are salt-sensitive (Remus et al., 2009, Evrin et al., 2009, Kang et al., 2014), single molecule analysis of the reaction reveals that nearly 50% of the singly-loaded helicases are salt-stable, suggesting that at this stage Mcm2-7 ring closure has occurred (Ticau et al., 2015). This discrepancy between bulk and single-molecule studies currently lacks explanation, although may arise from the relatively high protein concentrations used in the bulk assays and/or by their limited kinetic sensitivity.
After the first loading event, Cdc6 re-associates with the ORC•Mcm2-7 complex, forming a short-lived ORC•Cdc6•Mcm2-7 (OCM) intermediate that is competent for recruitment of a second Cdt1•Mcm2-7 heptamer (Fernandez-Cid et al., 2013, Ticau et al., 2015). At present, the mechanics by which a second hexamer is recruited are unclear. It has been proposed that loading of the second hexamer requires equivalent interactions as the first, with the Mcm3 C-terminus being required for both loading steps (Frigola et al., 2013). However, if loading the second helicase into a double hexamer requires an interaction with ORC•Cdc6, this mechanism must involve some fairly complicated acrobatics, mainly because the first loaded Mcm2-7 complex should still be present to block the MCM-binding surface on ORC, and because the second Mcm2-7 complex must be placed at a location more than ~100 Å from ORC’s initial binding site on DNA, and in an inverted orientation from the first helicase. This spatial problem could be solved if two ORCs were used for the loading of two helicases with opposing polarity; however, single-molecule analysis has demonstrated that a single, stably bound ORC is sufficient to load a double hexamer (Ticau et al., 2015). Because of these steric complications, an alternative mechanism has been proposed wherein the first, loaded hexamer templates loading of the second through a distinct mechanism that does not require direct ORC•Cdc6 interactions (Ticau et al., 2015). Consistent with this idea, the kinetics of loading the second hexamer are slower than the first (Fernandez-Cid et al., 2013), as is the release of Cdc6 and Cdt1 (Ticau et al., 2015), suggesting different reaction intermediates are accessed.
While the model described above does not directly account for the demonstrated role of the budding yeast Mcm3 WH domain in the loading of both the first and second Mcm2-7 hexamers (Frigola et al., 2013), previous work has revealed that deletion of the Mcm3 C-terminus results in a dramatic loss of Mcm2-7 ATPase activity (Sun et al., 2014). Since proper ATPase function by the Mcm2-7 hexamer is critical at all stages of the loading reaction in S. cerevisiae, this loss of activity in the Mcm3 WH domain mutant could account for the observed defect in loading the second hexamer. Accordingly, Mcm2-7 complexes with individual active site mutations also show defective recruitment (particularly Mcm2 Walker A and Mcm6 arginine finger mutants), Cdt1 release, and dramatically reduced levels of salt-stable double hexamers (Kang et al., 2014, Coster et al., 2014). Although a marginal level of double-hexamer formation can be achieved with particular MCM ATPase mutants, these loaded complexes are generally deficient for downstream activation events and cannot drive DNA replication in vitro (Kang et al., 2014). Collectively, the present data indicate that ATP binding and hydrolysis by S. cerevisiae Mcm2-7 is required from the early steps of helicase recruitment to the concluding activation events. These results are difficult to rationalize with the observation that Xenopus MCM6 and MCM7 Walker A mutants are loaded normally (Ying and Gautier, 2005), but this finding may again reflect species-specific differences.
Although double-hexamer loading is critically dependent on the ATPase activity of Mcm2-7, ATP hydrolysis by Cdc6 and ORC perform additional roles that are consistent with a requirement for their ATPase activity in vivo (Zwerschke et al., 1994, Weinreich et al., 1999, Wang et al., 1999, Schepers and Diffley, 2001, Takahashi et al., 2002, Bowers et al., 2004). ATP hydrolysis by Orc1 is dispensable for the loading reaction (Evrin et al., 2013, Fernandez-Cid et al., 2013, Coster et al., 2014), but resets the initiation machinery to allow reiterative loading of MCM double hexamers at origins (Bowers et al., 2004). By comparison, Cdc6 has been reported to function as a quality control factor, whose ATPase activity ensures that defective or incomplete Mcm2-7 complexes are released from origins (Frigola et al., 2013, Kang et al., 2014, Coster et al., 2014). Another major function of Cdc6 ATPase activity appears to be in the final step of pre-RC formation, where it is responsible for releasing successfully loaded double hexamers from ORC, an event that enables the downstream activation of the helicase (Chang et al., 2015). Loss of Cdc6 ATPase activity also results in inefficient Cdt1 release (Randell et al., 2006), although ATP turnover by Mcm2-7 appears primarily responsible for Cdt1 ejection (Kang et al., 2014, Coster et al., 2014). Overall, Cdc6 functions as a linchpin factor critical for multiple events in the loading process.
Helicase activation
Although a double MCM2-7 hexamer is loaded during pre-RC formation, at replication forks this structure is dissolved to yield two single MCM2-7 hexamers that move in opposite directions (Yardimci et al., 2010, Ticau et al., 2015). On their own, single MCM2-7 hexamers possess little or no helicase activity (Ishimi, 1997, Lee et al., 2000, Kaplan et al., 2003, You et al., 2003, Ilves et al., 2010). Initially, this observation was difficult to rationalize with the strong evidence that pointed towards a role for the complex as the replicative helicase, but it is now appreciated that MCM2-7 is incorporated into a larger Cdc45•MCM2-7•GINS (CMG) complex that possesses robust helicase activity. Like pre-RC assembly, double hexamer dissolution and CMG formation occur through a highly orchestrated and interconnected series of events that provide yet another means for regulating the DNA replication initiation reaction.
Identification of the Cdc45•MCM2-7•GINS (CMG) complex
The MCM2-7 genes are essential proteins in DNA replication (Moir et al., 1982, Maine et al., 1984, Gibson et al., 1990, Dalton and Whitbread, 1995) and their phylogenetic lineage to helicases led to early suggestions that these factors might function as nucleic acid motor proteins (Koonin, 1993). Together with work on archaeal MCMs, which have been found to possess robust helicase activity (Chong et al., 2000, Kelman et al., 1999, Shechter et al., 2000), these data indicated that the MCM2-7 complex likely served as the replicative helicase in eukaryotes. Interestingly, initial in vitro studies were unable to detect helicase activity from the eukaryotic MCM2-7 complex, although an MCM subcomplex composed of subunits (4,6,7)2 showed weak helicase activity (Ishimi, 1997, Lee et al., 2000, Kaplan et al., 2003, You et al., 2003). These data suggested that the MCM2-7 assembly might play a more limited role in replication, such as origin melting. However, the persistence of MCM2-7 at replication forks (Aparicio et al., 1997), coupled with the observed S-phase arrest after MCM2-7 inactivation (Labib et al., 2000, Pacek and Walter, 2004, Shechter et al., 2004), provided irrefutable evidence for the role of the complex in elongation (Todorov et al., 1995, Krude et al., 1996). Nevertheless, the demonstration of helicase activation remained elusive.
Eventually, conditions were identified that recovered in vitro helicase activity for budding yeast Mcm2-7 (Bochman and Schwacha, 2008). Roughly contemporaneously, several lines of analysis began to reveal that the helicase also existed within a larger macromolecular assembly (Masuda et al., 2003, Calzada et al., 2005, Pacek et al., 2006, Gambus et al., 2006, Moyer et al., 2006). Three parallel studies – two using an unbiased assessment of Cdc45 and GINS interacting factors and a third using a candidate-based approach to determine the composition of the helicase stalled at replication forks – have now demonstrated that MCM2-7 forms a stable complex with Cdc45 and GINS (Pacek et al., 2006, Gambus et al., 2006, Moyer et al., 2006). Cdc45 and GINS are essential replication fork components (Moir et al., 1982, Hopwood and Dalton, 1996, Dalton and Hopwood, 1997, Zou et al., 1997, Kamimura et al., 1998, Kanemaki et al., 2003, Masuda et al., 2003), which mimic the chromatin association pattern of MCM subunits (Takayama et al., 2003, Kubota et al., 2003). Each MCM2-7 binds a single copy of Cdc45 and one GINS tetramer to form a stable, eleven-protein complex termed the CMG (for Cdc45•MCM2-7•GINS). The CMG has been shown to possess robust helicase activity (Moyer et al., 2006, Ilves et al., 2010), and bioinformatics analyses have indicated that a full set of CMG factors are likely conserved from archaea to eukaryotes (Makarova et al., 2012), with recent work demonstrating that, as observed for the eukaryotic accessory factors, archaea’s Cdc45 and GINS homologs activate MCM in vitro and in vivo (Xu et al., 2016).
Regulating helicase activation
Appropriate recruitment of the factors necessary for CMG formation constitutes a highly-regulated step towards replisome formation. Consistent with the Cdc6-dependent release of the MCM2-7 double hexamer from ORC (Chang et al., 2015), helicase activation steps no longer require the function of either ORC or Cdc6 (Hua and Newport, 1998, Rowles et al., 1999, Jares and Blow, 2000, Walter, 2000, Yeeles et al., 2015). The newly formed MCM2-7 double hexamer serves itself as a binding platform and target for the Dbf4-dependent kinase (DDK) that, in collaboration with S-phase cyclin-dependent kinase (S-CDK), constitute the minimal set of kinases needed to regulate the recruitment of Cdc45 and GINS. Together, the activity of these two kinases coordinate the assembly of a pre-IC that is poised for activation and formation of a bidirectional replication fork (Figure 6) (Aladjem, 2007).
DDK is formed by the direct association of the Cdc7 kinase with Dbf4, a protein co-factor that relieves Cdc7 autoinhibition (Kitada et al., 1992, Jackson et al., 1993, Dowell et al., 1994) and that is regulated in a cell-cycle dependent fashion (Cheng et al., 1999, Oshiro et al., 1999, Nougarede et al., 2000). DDK phosphorylates multiple MCM subunits (Lei et al., 1997, Weinreich and Stillman, 1999, Masai et al., 2006, Montagnoli et al., 2006, Cho et al., 2006, Sheu and Stillman, 2006, Tsuji et al., 2006) and is required for initiating DNA replication (Chapman and Johnston, 1989, Hollingsworth and Sclafani, 1990, Jiang et al., 1999a). Like DDK, S-CDK is required for replication (Broek et al., 1991, Hayles et al., 1994) and exists in an inactive conformation by engaging Sic1, a repressive protein that restricts S-CDK activity into late G1/S phase (Mendenhall, 1993, Donovan et al., 1994, Schwob et al., 1994, Knapp et al., 1996, Schneider et al., 1996). Redundant regulation of S-CDK occurs by the cell-cycle dependent expression of activating cyclins (Abreu et al., 2013). It is currently unclear whether there is a conserved sequential order for DDK and CDK activity. In a X. laevis egg extract system, DDK activity is independent of CDK, and exposing the pre-RC first to CDK fully ablates DNA replication initiation (Jares and Blow, 2000, Walter, 2000), suggesting that DDK acts first. Such a consensus has not been found with S. cerevisiae proteins, where the sequential action of DDK and CDK depends on the experimental setup (Nougarede et al., 2000, Heller et al., 2011) and has been found to be inconsequential in a fully in vitro reconstituted DNA replication system (Yeeles et al., 2015).
DDK is recruited to Mcm2-7 through interactions with Mcm4 and Mcm2 (Sheu and Stillman, 2010, Ramer et al., 2013) and targets specific MCM subunits within the context of a double, but not a single, MCM2-7 hexamer (Francis et al., 2009, Evrin et al., 2014, Sun et al., 2014, Kang et al., 2014, Costa et al., 2014). Notably, the MCM subunits responsible for recruiting DDK are non-adjacent within the context of a single hexamer; however, the rotational offset within the double hexamer results in the adjacent placement of MCM2 and MCM4 from separate hexamers (Costa et al., 2014, Sun et al., 2014, Li et al., 2015), thus forming a composite DDK-interaction interface. In addition, MCM2, MCM4, and MCM6, which are all phosphorylated by DDK (Lei et al., 1997, Masai et al., 2006, Tsuji et al., 2006, Sheu and Stillman, 2006, Cho et al., 2006, Bruck and Kaplan, 2009, Randell et al., 2010), reside in close proximity within the double hexamer and the interface between the two hexamers is somewhat splayed apart at this position, creating a gap that may provide DDK access to the N-terminal serine/threonine rich domains (NSD) of its MCM targets (Sheu and Stillman, 2010, Sun et al., 2014, Li et al., 2015). Collectively, the available data suggest that the DDK-dependent activation of MCM2-7 is temporally regulated by selective recruitment of the kinase to and action upon a double hexameric MCM2-7, a structure that uniquely positions the relevant MCM subunits in close proximity (Sun et al., 2014). In addition to DDK, budding yeast Mcm2-7 has been shown to be phosphorylated by CDK and Mec1 kinases, which sensitize Mcm2-7 to DDK activity (Devault et al., 2008, Randell et al., 2010); however, in vitro DDK alone is required for activation of S. cerevisiae Mcm2-7 (Yeeles et al., 2015).
In S. cerevisiae, the DDK-dependent phosphorylation of Mcm2-7 is important for facilitating the Sld3 and Sld7-chaperoned recruitment of Cdc45 onto the double hexamer (Masai et al., 2006, Yabuuchi et al., 2006, Sheu and Stillman, 2006, Tanaka et al., 2011a). In humans, two proteins known as Treslin and MTBP (for MDM2 binding protein) have been proposed to represent functional homologs of budding yeast Sld3 and Sld7, respectively (Matsuno et al., 2006, Sanchez-Pulido et al., 2010, Sansam et al., 2010, Kumagai et al., 2010, Bruck and Kaplan, 2015c, Boos et al., 2013). In vitro, Cdc45 recruitment depends on the prior association of Sld3 and Sld7 with the DDK-phosphorylated N-termini of Mcm2, Mcm4, and Mcm6 (Deegan et al., 2016, Fang et al., 2016); however, in vivo Sld3/7 and Cdc45 form a stable complex and may be co-recruited to double hexamers (Kamimura et al., 2001, Tanaka et al., 2011a, Tanaka et al., 2011b). Notably, in the absence of DDK, no additional initiation factors are recruited to Mcm2-7 double hexamers. Conversely, in the absence of CDK activity, Cdc45 as well as Sld3 and Sld7 are still efficiently recruited (Heller et al., 2011, Yeeles et al., 2015). Thus, recruitment of Cdc45 appears to rely exclusively on DDK.
How does DDK-dependent MCM2-7 phosphorylation facilitate Cdc45 recruitment? Although we lack a complete mechanistic picture for DDK-dependent activation, three distinct MCM modifications have been identified that result in DDK bypass, suggesting there exists a complex role for DDK that likely involves a concerted change within multiple MCM subunits. First, hyperphosphorylation of the S. cerevisiae Mcm4 N-terminus relieves an autoinhibitory function that this region imposes on the helicase (indeed, deletion of the Mcm4 N-terminus bypasses the DDK requirement) (Sheu and Stillman, 2006, Sheu and Stillman, 2010). Although a mechanism entailing a steric block to the Cdc45 binding site or the adoption of a different non-permissive orientation would satisfy this observation, within the atomic structure of the S. cerevisiae Mcm2-7 double hexamer the Mcm4 N-terminus is disordered, suggesting that autoinhibition is relieved by an alternative mechanism (Li et al., 2015). Second, phosphorylation of Mcm4 and Mcm6 generates a phosphopeptide binding site for recruiting Sld3 (phosphomimetic mutants of these MCM subunits can bypass the requirement for DDK) (Deegan et al., 2016). Finally, a mutation within budding yeast Mcm5, known as the bob1 allele, has been found to bypass DDK-activated initiation (Hardy et al., 1997).
In addition to DDK-dependent recruitment of Cdc45, Mcm2-7 activation requires an association with GINS and, for origin firing, with Replication Protein A (RPA), MCM10, Polα, Polε, and CTF4 (van Deursen et al., 2012, Yeeles et al., 2015). Notably, CDK’s action on the growing, chromatin-associated replication assembly, as well as on replicative proteins yet to be integrated into the replisome, results in the recruitment of the remaining initiation factors that are needed for origin firing. In this swarm of activity, CDK targets Cdc45, Sld2, Sld3, and Sld7 (Masumoto et al., 2002, Zegerman and Diffley, 2007, Tanaka et al., 2007, Muramatsu et al., 2010, Heller et al., 2011, Yeeles et al., 2015), and, together with a protein known as Dpb11, facilitates recruitment of a multiprotein assembly composed of GINS, Sld2, and Polε to the growing complex (Araki et al., 1995, Kamimura et al., 1998, Kamimura et al., 2001, Takayama et al., 2003, Muramatsu et al., 2010, Tanaka et al., 2013). Thus, DDK and CDK drive formation of an active helicase by phosphorylating key assembly factors that allow for the sequential recruitment of Cdc45 and GINS. For its part, MCM10 is recruited by interactions with the CMG (Homesley et al., 2000, Wohlschlegel et al., 2002, Douglas and Diffley, 2016) and further appears to help stabilize the replisome once formed (Gregan et al., 2003, Ricke and Bielinsky, 2004). Polε is likewise recruited directly to the CMG by an interaction with GINS (Araki et al., 1995, Muramatsu et al., 2010), whereas Polα is physically coupled to the helicase by CTF4 (Villa et al., 2016, Simon et al., 2014).
Many questions remain for the events that transition the growing MCM2-7 complex into a competent bi-directional replication fork. For example, it is currently unclear at what point and how origin melting occurs, or how an MCM2-7 double hexamer, which encircles duplex DNA, transitions into two uncoupled CMG particles that encircle single-stranded DNA (Yardimci et al., 2010, Ticau et al., 2015). Interestingly, recent data suggests that the process of origin melting and strand extrusion may be interdependent and temporally intertwined with CMG assembly. Indeed, S. cerevisiae Sld2, Sld3, and Dpb11, which chaperone GINS into a complex with the helicase, compete with GINS for Mcm2-7 binding, a function that is relieved in the presence of single-stranded DNA (Bruck et al., 2011, Bruck and Kaplan, 2011, Bruck and Kaplan, 2014, Bruck and Kaplan, 2015c, Dhingra et al., 2015) (the metazoan Sld2 and Sld3 equivalents also bind single-stranded DNA (Sangrithi et al., 2005, Ohlenschlager et al., 2012, Bruck and Kaplan, 2015c)). Thus, GINS binding may not only be anticipated by origin melting and the generation of single-stranded DNA, but by strand extrusion as well, as the separated origin strands may be contained within the Mcm2-7 double hexamer for a period of time (Sun et al., 2014, Li et al., 2015). The topological transformations to DNA in the context of the maturing CMG complex are likely aided in by the phosphorylation of Mcm2 by DDK (Lei et al., 1997, Bruck and Kaplan, 2009), an event that weakens the Mcm2/5 gate and promotes ring opening (Bruck and Kaplan, 2015a, Bruck and Kaplan, 2015b). Consistent with this idea, origin melting and strand extrusion have been reported to occur concomitantly with DDK-dependent activation but prior to complete CMG assembly (Bruck and Kaplan, 2015a). Interestingly, recent data suggests that the single-stranded DNA binding activity of budding yeast Mcm10 is required for initiation and that it may help stabilize the origin melting reaction (Perez-Arnaiz and Kaplan, 2016).
Despite the available insights, the mechanisms that transition a loaded double hexamer into an active, bi-directional replication fork remain shrouded in mystery. An important future direction will be to understand the timing of different events in the lifecycle of the helicase, such as when double-hexamer dissociation and replisome factor recruitment occurs (such as MCM10, Polα and RPA), as compared to origin melting and strand extrusion. The physical mechanisms that coordinate CMG structural changes with origin remodelling likewise remain ill-defined, and require structural studies of different intermediates that, at present, have not been stably isolated.
Mechanism of CMG function
The cracked ring architecture of the MCM2-7 complex is an asset when it comes to loading the helicase around double-stranded DNA, but would seem to be a potential liability in terms of processive DNA unwinding functions (Bochman and Schwacha, 2008, Ilves et al., 2010, Petojevic et al., 2015). Although nucleotide binding by MCM2-7 helps to overcome a natural tendency of the helicase to splay open (Samel et al., 2014), cracked-ring conformations still readily form (Costa et al., 2011, Lyubimov et al., 2012), providing a physical basis for the weak in vitro helicase activity of MCM2-7 (Bochman and Schwacha, 2008, Ilves et al., 2010). Incorporation of MCM2-7 into the CMG counteracts the functional deficiencies of the MCM2-7 complex for DNA unwinding. 3D electron microscopic reconstructions of both the Drosophila and budding yeast CMG have revealed that GINS and Cdc45 seal off the MCM2/MCM5 discontinuity and reduce the conformational dynamics of the MCM2/5 gate, likely as a means to favor productive ATPase interface contacts (Figure 7A-B) (Costa et al., 2011, Costa et al., 2014, Abid Ali et al., 2016, Yuan et al., 2016). Thus, instead of being directly involved in translocation per se, it would seem that GINS and Cdc45 help shift the structural disposition of the MCM2-7 AAA+ domains into productive conformations. Consistent with this notion, incorporation of Drosophila MCM2-7 into the CMG increases the vmax of ATP hydrolysis by over 300-fold and results in a 10-fold higher affinity for DNA, resulting in a dramatic increase in helicase activity (Ilves et al., 2010). Interestingly, in addition to participating in the control of MCM2-7 ring status, both GINS (Boskovic et al., 2007, Ilves et al., 2010) and Cdc45 (Krastanova et al., 2012, Bruck and Kaplan, 2013, Szambowska et al., 2014) have been shown to bind DNA. Cdc45, which is a catalytically-defunct homolog of the RecJ exonuclease (Sanchez-Pulido and Ponting, 2011, Simon et al., 2016), also appears to assist with capture of the leading strand during transient CMG gate opening (Petojevic et al., 2015), thereby possibly playing a protective role during replication fork stalling (Bruck and Kaplan, 2013).
Figure 7.
CMG helicase organization and dynamics. (A) MCM2-7 adopts a spiral, cracked-ring architecture with a discontinuity between MCM2 and MCM5. Upon incorporation into the CMG, Cdc45 and GINS convert the helicase into a planar form and seal off the MCM2/5 gate (EM density maps = EMD-1835 and EMD-1833 for MCM2-7 and CMG, respectively). (B) A view of the CMG from the AAA+ face illustrating the overall architecture of the complex (PDB = 3JC5 (Yuan et al., 2016)). (C) At least two conformational states exist for the CMG, a constricted state in which the AAA+ and NTD rings are coplanar (bottom panel), and a relaxed state where one end of the AAA+ tier lifts up from NTD tier (top panel). These conformations appear coupled to alterations in the relative disposition of the gating subunits, MCM2 and MCM5 (constricted and relaxed conformer PDB codes = 3JC5 and 3JC7, respectively). (D) The activated CMG helicase is thought to translocate along single-stranded DNA, unwinding downstream template through a combined steric exclusion and DNA wrapping mechanism. A color version of this figure is available online.
Structural analysis of the CMG reveals the presence of two distinct conformers that are, as observed for the MCM2-7 helicase core, differentiated by the relative positioning of the gating subunits (Yuan et al., 2016, Abid Ali et al., 2016). As the MCM NTD is stabilized by extensive contacts with GINS and Cdc45, changes within the AAA+ ATPase ring appear primarily responsible for the alterations observed between the two conformational states (Costa et al., 2011, Costa et al., 2014, Abid Ali et al., 2016, Yuan et al., 2016). Altering the position of the MCM2/5 gate appears to propagate structural changes around the MCM2-7 ring that result in significant translational and rotational rearrangement of MCM6 and, to a lesser extent, MCM4. These changes in turn are coupled to a transition from a constricted conformer state, where the MCM2-7 NTD and CTD tiers are co-planar, to a relaxed, tilted form, which generates an asymmetric expansion of the gap between the NTD and CTD tiers on one side of the ring (Figure 7C). It has been suggested that switching between these conformations could elicit a pumpjack motion, with the CTD tier rocking with respect to a stable NTD tier during DNA translocation (Yuan et al., 2016), and that these dynamics could be used to drive a linear, as opposed to rotational, mechanism of DNA translocation (Yuan et al., 2016, Abid Ali et al., 2016). Higher-resolution structures of distinct translocation intermediates, coupled with single-molecule measurements of ring dynamics and displacement, will be needed to test these ideas.
An interesting point of discussion in the field has been whether the ATPase regions of MCMs serve as single or double-stranded DNA translocases. Based on a distant phylogenetic kinship to RuvB, a known double-strand translocase, MCM2-7 was initially proposed to move along duplex DNA, with the NTD-associated GINS/Cdc45 accessory subunits serving as a ploughshare to separate the two strands (Takahashi et al., 2005). However, subsequent biochemical studies of the Drosophila CMG have shown that the complex requires a fork for unwinding (Ilves et al., 2010), consistent with both archaeal MCM and metazoan MCM4,6,7 functioning as single-stranded DNA translocases on model substrates (Ishimi, 1997, Kelman et al., 1999, Shechter et al., 2000, Moyer et al., 2006, Kang et al., 2012). Along these lines, studies in archaea have reported that an MCM construct lacking the NTD tier (and that has no associated GINS/Cdc45) still possesses helicase activity (Barry et al., 2007). Moreover, replisomes have been challenged with leading and lagging strand roadblocks in an X. laevis extract system, with only leading strand roadblocks proving capable of stalling fork progression (MCM2-7 is a 3′→5′ helicase that tracks on the leading strand) (Fu et al., 2011). Consistent with this observation, recent analysis of the pattern of forked DNA crosslinking to the CMG reveals that the leading strand shows robust crosslinking to all MCM2-7 subunits, as well as Cdc45, while lagging strand crosslinks are markedly fewer and those that occur are all on the exterior surface of MCM2-7 (Petojevic et al., 2015). Collectively, these data strongly indicate that the CMG functions as a single-stranded DNA translocase.
Structural analysis of the apo and substrate-bound CMG states also support a single-strand tracking model. In particular, upon binding of the CMG to a forked duplex substrate, nucleic acid density can be visualized within the central channel whose dimensions appear sufficient to accommodate single but not double-stranded DNA (Abid Ali et al., 2016). This finding is consistent with a previous structural analysis of the CMG bound to duplex DNA containing a single 3′-tail, which showed that the duplex region of a DNA fork is positioned outside the central channel in an orientation consistent with the known 3′→5′ polarity of DNA engagement (Costa et al., 2014). Interestingly, the high-resolution structure of the S. cerevisiae CMG reveals that the WH domains of Mcm5 and Mcm6, which are disordered within the context of the double hexamer (Li et al., 2015), become ordered and take up a position within the central pore of the CMG that constricts the AAA+ pore to a point where it is too small to accommodate double-stranded DNA (Yuan et al., 2016). Overall, these lines of evidence largely corroborate the view that the CMG tracks along single-stranded templates, albeit with a potential capacity to switch to a double-stranded DNA mode of translocation, as might occur during replication termination, when forks converge (Dewar et al., 2015) (parenthetically, there is evidence that the E. coli DnaB helicase can undergo a transition between single and double-stranded translocation modes (Kaplan, 2000, Kaplan and O’Donnell, 2002)). Understanding the role of the WH domains awaits determination of substrate-bound CMG structures in which the disposition of the nucleic acid substrate and the WH domains are fully defined.
Overall, while a variety of mechanistic models exist to account for hexameric helicase activity, the so-called steric exclusion model appears most compatible with the observed data to date for the CMG. The steric exclusion model posits that translocation occurs along single-stranded DNA with the non-template strand being excluded from the central channel. One assumption of this framework has been that the lagging strand is functionally passive, implying that once displaced, it is inconsequential to helicase function. However, emerging lines of evidence are starting to suggest that the lagging strand may wrap around the external surface of the CMG and support helicase activity (Figure 7D). Consistent with this notion, DNA interactions with the external surface of archaeal MCM is integral to helicase function (Rothenberg et al., 2007, Costa et al., 2008, Graham et al., 2011, Graham et al., 2016). DNA binding to the MCM external surface appears at least partially facilitated by the NTD-A region, which can undergo large structural rearrangements to expose a DNA-binding site (Fletcher et al., 2003, Chen et al., 2005, Costa et al., 2008, Miller et al., 2014). Whether the eukaryotic MCM NTD-A also facilitates DNA binding is unclear; however, the MCM AAA+ domain contains a conserved, surface-exposed β–hairpin that has been reported to directly engage DNA (Brewster et al., 2010, Graham et al., 2011) and, in the context of the eukaryotic CMG, is integral to helicase function (Petojevic et al., 2015). Thus, the CMG may function by a mechanism of single-stranded DNA translocation that results in both steric exclusion and wrapping of lagging strand DNA (Graham et al., 2011).
Regulatory mechanisms for preventing re-initiation
Origin licensing and firing are the most tightly regulated events in the process of DNA replication initiation. Multiple interdependent and redundant mechanisms ensure that origins can be marked and utilized only within a certain window of the cell cycle. While the previous discussion of helicase activation has demonstrated the role of CDK and DDK in positively regulating initiation, a plethora of mechanisms exist to negatively regulate replication licensing and prevent re-replication as well. This section will discuss some general principles for preventing re-replication; although regulatory approaches show substantial evolutionary diversification and expansion, certain common themes do exist. For a more thorough discussion of the regulation of DNA replication initiation in eukaryotes, we point the reader to a number of dedicated reviews on the topic (see (Arias and Walter, 2007, Hook et al., 2007, Masai et al., 2010, Araki, 2010, Siddiqui et al., 2013)).
Initiator and helicase regulation
By temporally separating helicase loading from activation, S-CDK can simultaneously activate the helicase while also inhibiting helicase loading factors to prevent origin re-firing. A major means by which CDK protects against re-licensing in S-phase is to regulate initiator interactions and their association with chromatin. Although budding yeast ORC remains stably bound throughout the cell cycle (Diffley et al., 1994, Liang and Stillman, 1997, Fujita et al., 1998), vertebrate Orc1 is the target of CDK-dependent and CDK-independent mechanisms that remove it from chromatin (Figure 8) (Rowles et al., 1999, Findeisen et al., 1999, Natale et al., 2000, Li and DePamphilis, 2002, Li et al., 2004, Kreitz et al., 2001, Sun et al., 2002). In both yeast and metazoans, initiation factors are targeted for proteolysis in a CDK-dependent fashion, with budding yeast exclusively targeting Cdc6 (Piatti et al., 1995, Drury et al., 1997, Perkins et al., 2001, Mimura et al., 2004) and metazoans targeting both Orc1 and Cdc6 (Mendez et al., 2002, Tatsumi et al., 2003, Ohta et al., 2003, Lidonnici et al., 2004, Kalfalah et al., 2015). In S. pombe, Orc2 is also targeted by CDK to prevent re-licensing, but the mechanism utilized in this instance is unclear (Vas et al., 2001, Wuarin et al., 2002). In S. cerevisiae, CDK-dependent control is further extended to regulate Cdc6 transcription and nuclear localization (Moll et al., 1991, Honey and Futcher, 2007); metazoan Cdc6 likewise undergoes CDK-dependent nuclear export to limit replication (Saha et al., 1998, Jiang et al., 1999b, Petersen et al., 1999). Interestingly, budding yeast Orc6 and Cdc6 stably associate with CDK during S-phase and this interaction sterically inhibits interactions necessary for pre-RC assembly, such as with Cdt1 (Mimura et al., 2004, Wilmes et al., 2004, Chen and Bell, 2011). In metazoans, Cdt1 has not yet been reported to interact with Orc6.
Figure 8.
Mechanisms to prevent re-replication. Multiple, redundant mechanisms are utilized to prevent re-licensing of origins after S-phase has initiated. Whereas yeast seem to exclusively utilize CDK-dependent mechanisms to prevent re-licensing, metazoans also employ CDK-independent pathways for negatively regulating Cdt1 activity. A color version of this figure is available online.
In addition to regulating ORC and Cdc6, in S. cerevisiae CDK targets Mcm2-7 to prevent new, productive interactions with chromatin-bound ORC•Cdc6 complexes. This action occurs through the CDK-dependent nuclear exclusion and export of Mcm2-7 (Labib et al., 1999, Nguyen et al., 2001, Tanaka and Diffley, 2002, Liku et al., 2005). There is currently no evidence that CDK targets metazoan MCM2-7 directly, although phosphorylation of the initiator reduces helicase association with origins (Findeisen et al., 1999). Interestingly, a new, CDK-independent MCM regulatory mechanism has recently been identified in budding yeast that involves the SUMOylation of all members of the Mcm2-7 hexamer in G1. The SUMO modification in turn, through phosphatase recruitment, prevents phosphorylation-dependent helicase activation and thus negatively regulates initiation (Wei and Zhao, 2016). How the SUMO pathway integrates with the other mechanisms that regulate initiation is currently unclear.
Cdt1: a master regulatory nexus
Cdt1 plays an essential part in the loading of MCM2-7 onto DNA as a stable double hexamer. As such, Cdt1 turns out to be a common regulatory point. S. cerevisiae Cdt1 activity is restricted by the CDK-dependent inhibition of its interaction with Orc6 (Chen and Bell, 2011) and by nuclear export of Cdt1 in G1 (Tanaka and Diffley, 2002). Conversely, metazoan and S. pombe Cdt1 protein levels are regulated in a cell-cycle dependent fashion such that Cdt1 is actively degraded upon S-phase entry (Figure 8) (Wohlschlegel et al., 2000, Nishitani et al., 2000, Gopalakrishnan et al., 2001, Nishitani et al., 2001, Zhong et al., 2003, Nishitani et al., 2004). Degradation of Cdt1 is restricted to S-phase by an interaction with chromatin-bound PCNA, which serves as a platform for Cdt1 recognition and ubiquitination by the Cullin-RING ligase 4 (CRL4) ubiquitin ligase (Arias and Walter, 2005, Arias and Walter, 2006, Senga et al., 2006, Hu and Xiong, 2006, Guarino et al., 2011). A conserved PCNA-interacting protein (PIP) degron within the Cdt1 N-terminus facilitates the interaction with chromatin-bound PCNA and is required for degradation (Senga et al., 2006, Havens and Walter, 2009, Havens and Walter, 2011). Similarly, the CDK-dependent phosphorylation of Cdt1 leads to its recognition and ubiquitination by the SCFSkp2 E3 ubiquitin ligase, providing an additional mechanism to limit Cdt1 protein levels and prevent re-replication (Li et al., 2003, Liu et al., 2004, Sugimoto et al., 2004, Thomer et al., 2004, Kondo et al., 2004, Nishitani et al., 2006). Notably, Cdt1 ubiquitination can be reversed through the function of the ubiquitin hydrolase USP37, whose activity stabilizes Cdt1 and promotes helicase loading (Hernandez-Perez et al., 2016).
Metazoan Cdt1 is also uniquely regulated by binding to a partner protein, Geminin. Geminin is a coiled-coil protein (Saxena et al., 2004, Lee et al., 2004) that was initially identified in Xenopus egg extract screens for proteins destabilized in mitosis (McGarry and Kirschner, 1998). This study, along with many others, revealed that Geminin, a nuclear protein whose levels become elevated during S-phase, targets and restricts Cdt1 activity to G1, thus preventing Cdt1-induced re-replication (Quinn et al., 2001, Tada et al., 2001, Mihaylov et al., 2002, Cook et al., 2004, Yoshida et al., 2005, Lutzmann et al., 2006, Kerns et al., 2007). Geminin binds directly to Cdt1 (Wohlschlegel et al., 2000, Lutzmann et al., 2006, De Marco et al., 2009), and this interaction has been shown to inhibit Cdt1 binding to mouse MCM6 (Yanagi et al., 2002). Recent work demonstrates that Geminin binding to Cdt1 inhibits ORC-dependent helicase loading but not recruitment (Wu et al., 2014a), a finding consistent with studies in budding yeast showing Cdt1-independent recruitment of the helicase to the origin-bound initiator (Frigola et al., 2013, Fernandez-Cid et al., 2013).
Interestingly, Geminin serves two seemingly dichotomous functions. Although clearly an inhibitor of licensing, Geminin is also required for replication initiation through its ability to stabilize Cdt1 levels by protecting the protein from degradation (Ballabeni et al., 2004, Narasimhachar and Coue, 2009). Thus, metazoans regulate Geminin protein levels and cellular localization to liberate Cdt1 from Geminin and promote pre-RC assembly in G1 (McGarry and Kirschner, 1998, Tsunematsu et al., 2013, Dimaki et al., 2013). In addition, the stoichiometry of the Cdt1•Geminin complex has been suggested to regulate Geminin’s activity, such that a lower order Cdt1:Geminin complex (1:2) is permissive to pre-RC assembly whereas a higher order (2:4) complex is not (De Marco et al., 2009). How this stoichiometry is regulated in the context of the other regulatory mechanisms governing Geminin and Cdt1 protein levels and cellular localization is not known.
Concluding remarks
At this point, many of the major events and intermediates that facilitate pre-RC and pre-IC assembly have been defined. However, between these stable intermediates there exist multiple, undefined dynamic and transient protein interactions, modifications, and exchanges that represent critical steps towards building a replisome. The questions that remain are both broad and specific: how does ORC balance sequence preference and trans-acting chromatin contextual cues when selecting origins? What is the role of Cdt1 in MCM2-7 loading and, given the minimal conservation of this protein, is Cdt1 mechanism conserved across eukarya? Does MCM2-7 melt origins and unwind duplex DNA (as part of the CMG) using an overlapping or mutually exclusive set of protein-nucleic acid interactions, and how is this functional switch regulated? What type of molecular dance must take place to transition the pre-RC product, the double-stranded DNA-bound MCM2-7 double hexamer, into the CMG, and then again into a bi-directional replication fork defined by two single helicases bound to single-stranded DNA? It will be critical to address these and other questions in multiple model eukaryotic organisms to begin to understand what level of conservation can be expected for such a complex process. Future work that aims to resolve these and other questions will undoubtedly reveal exciting new mechanisms that underlie the highly orchestrated and regulated process of replication initiation in eukarya.
Acknowledgments
The authors acknowledge Berger and Botchan lab members past and present for valuable discussion and helpful advice in preparing this review.
Footnotes
Disclosure statement
This work was supported by an NIH NRSA postdoctoral fellowship (F32GM116393, to MWP) and by the NCI (R01CA030490, to JMB and MRB).
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]
- Abreu CM, Kumar R, Hamilton D, Dawdy AW, Creavin K, Eivers S, Finn K, Balsbaugh JL, O’connor R, Kiely PA, Shabanowitz J, Hunt DF, Grenon M, Lowndes NF. Site-specific phosphorylation of the DNA damage response mediator rad9 by cyclin-dependent kinases regulates activation of checkpoint kinase 1. PLoS Genet. 2013;9:e1003310. doi: 10.1371/journal.pgen.1003310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adachi Y, Usukura J, Yanagida M. A globular complex formation by Nda1 and the other five members of the MCM protein family in fission yeast. Genes Cells. 1997;2:467–79. doi: 10.1046/j.1365-2443.1997.1350333.x. [DOI] [PubMed] [Google Scholar]
- Aggarwal BD, Calvi BR. Chromatin regulates origin activity in Drosophila follicle cells. Nature. 2004;430:372–6. doi: 10.1038/nature02694. [DOI] [PubMed] [Google Scholar]
- Aladjem MI. Replication in context: dynamic regulation of DNA replication patterns in metazoans. Nat Rev Genet. 2007;8:588–600. doi: 10.1038/nrg2143. [DOI] [PubMed] [Google Scholar]
- Aparicio OM, Weinstein DM, Bell SP. Components and dynamics of DNA replication complexes in S. cerevisiae: redistribution of MCM proteins and Cdc45p during S phase. Cell. 1997;91:59–69. doi: 10.1016/s0092-8674(01)80009-x. [DOI] [PubMed] [Google Scholar]
- Araki H. Cyclin-dependent kinase-dependent initiation of chromosomal DNA replication. Curr Opin Cell Biol. 2010;22:766–71. doi: 10.1016/j.ceb.2010.07.015. [DOI] [PubMed] [Google Scholar]
- Araki H, Leem SH, Phongdara A, Sugino A. Dpb11, which interacts with DNA polymerase II(epsilon) in Saccharomyces cerevisiae, has a dual role in S-phase progression and at a cell cycle checkpoint. Proc Natl Acad Sci U S A. 1995;92:11791–5. doi: 10.1073/pnas.92.25.11791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aravind L, Landsman D. AT-hook motifs identified in a wide variety of DNA-binding proteins. Nucleic Acids Res. 1998;26:4413–21. doi: 10.1093/nar/26.19.4413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arias EE, Walter JC. Replication-dependent destruction of Cdt1 limits DNA replication to a single round per cell cycle in Xenopus egg extracts. Genes Dev. 2005;19:114–26. doi: 10.1101/gad.1255805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arias EE, Walter JC. PCNA functions as a molecular platform to trigger Cdt1 destruction and prevent re-replication. Nat Cell Biol. 2006;8:84–90. doi: 10.1038/ncb1346. [DOI] [PubMed] [Google Scholar]
- Arias EE, Walter JC. Strength in numbers: preventing rereplication via multiple mechanisms in eukaryotic cells. Genes Dev. 2007;21:497–518. doi: 10.1101/gad.1508907. [DOI] [PubMed] [Google Scholar]
- Asano T, Makise M, Takehara M, Mizushima T. Interaction between ORC and Cdt1p of Saccharomyces cerevisiae. FEMS Yeast Res. 2007;7:1256–62. doi: 10.1111/j.1567-1364.2007.00299.x. [DOI] [PubMed] [Google Scholar]
- Atanasiu C, Deng Z, Wiedmer A, Norseen J, Lieberman PM. ORC binding to TRF2 stimulates OriP replication. EMBO Rep. 2006;7:716–21. doi: 10.1038/sj.embor.7400730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aves SJ, Liu Y, Richards TA. Evolutionary diversification of eukaryotic DNA replication machinery. Subcell Biochem. 2012;62:19–35. doi: 10.1007/978-94-007-4572-8_2. [DOI] [PubMed] [Google Scholar]
- Bae B, Chen YH, Costa A, Onesti S, Brunzelle JS, Lin Y, Cann IK, Nair SK. Insights into the architecture of the replicative helicase from the structure of an archaeal MCM homolog. Structure. 2009;17:211–22. doi: 10.1016/j.str.2008.11.010. [DOI] [PubMed] [Google Scholar]
- Balasov M, Huijbregts RP, Chesnokov I. Role of the Orc6 protein in origin recognition complex-dependent DNA binding and replication in Drosophila melanogaster. Mol Cell Biol. 2007;27:3143–53. doi: 10.1128/MCB.02382-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ballabeni A, Melixetian M, Zamponi R, Masiero L, Marinoni F, Helin K. Human geminin promotes pre-RC formation and DNA replication by stabilizing CDT1 in mitosis. EMBO J. 2004;23:3122–32. doi: 10.1038/sj.emboj.7600314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barry ER, Bell SD. DNA replication in the archaea. Microbiol Mol Biol Rev. 2006;70:876–87. doi: 10.1128/MMBR.00029-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barry ER, Lovett JE, Costa A, Lea SM, Bell SD. Intersubunit allosteric communication mediated by a conserved loop in the MCM helicase. Proc Natl Acad Sci U S A. 2009;106:1051–6. doi: 10.1073/pnas.0809192106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barry ER, Mcgeoch AT, Kelman Z, Bell SD. Archaeal MCM has separable processivity, substrate choice and helicase domains. Nucleic Acids Res. 2007;35:988–98. doi: 10.1093/nar/gkl1117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bell SP, Kobayashi R, Stillman B. Yeast origin recognition complex functions in transcription silencing and DNA replication. Science. 1993;262:1844–9. doi: 10.1126/science.8266072. [DOI] [PubMed] [Google Scholar]
- Bell SP, Mitchell J, Leber J, Kobayashi R, Stillman B. The multidomain structure of Orc1p reveals similarity to regulators of DNA replication and transcriptional silencing. Cell. 1995;83:563–8. doi: 10.1016/0092-8674(95)90096-9. [DOI] [PubMed] [Google Scholar]
- Bell SP, Stillman B. ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex. Nature. 1992;357:128–34. doi: 10.1038/357128a0. [DOI] [PubMed] [Google Scholar]
- Berbenetz NM, Nislow C, Brown GW. Diversity of eukaryotic DNA replication origins revealed by genome-wide analysis of chromatin structure. PLoS Genet. 2010;6:e1001092. doi: 10.1371/journal.pgen.1001092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bianconi E, Piovesan A, Facchin F, Beraudi A, Casadei R, Frabetti F, Vitale L, Pelleri MC, Tassani S, Piva F, Perez-Amodio S, Strippoli P, Canaider S. An estimation of the number of cells in the human body. Ann Hum Biol. 2013;40:463–71. doi: 10.3109/03014460.2013.807878. [DOI] [PubMed] [Google Scholar]
- Bleichert F, Balasov M, Chesnokov I, Nogales E, Botchan MR, Berger JM. A Meier-Gorlin syndrome mutation in a conserved C-terminal helix of Orc6 impedes origin recognition complex formation. Elife. 2013;2:e00882. doi: 10.7554/eLife.00882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bleichert F, Botchan MR, Berger JM. Crystal structure of the eukaryotic origin recognition complex. Nature. 2015;519:321–6. doi: 10.1038/nature14239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blow JJ. Preventing re-replication of DNA in a single cell cycle: evidence for a replication licensing factor. J Cell Biol. 1993;122:993–1002. doi: 10.1083/jcb.122.5.993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blow JJ, Laskey RA. A role for the nuclear envelope in controlling DNA replication within the cell cycle. Nature. 1988;332:546–8. doi: 10.1038/332546a0. [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;28:5865–73. doi: 10.1128/MCB.00161-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bochman ML, Schwacha A. The Mcm2–7 complex has in vitro helicase activity. Mol Cell. 2008;31:287–93. doi: 10.1016/j.molcel.2008.05.020. [DOI] [PubMed] [Google Scholar]
- Bochman ML, Schwacha A. DNA replication: Strand separation unravelled. Nature. 2015;524:166–7. doi: 10.1038/nature14643. [DOI] [PubMed] [Google Scholar]
- Boos D, Yekezare M, Diffley JF. Identification of a heteromeric complex that promotes DNA replication origin firing in human cells. Science. 2013;340:981–4. doi: 10.1126/science.1237448. [DOI] [PubMed] [Google Scholar]
- Bosco G, Du W, Orr-Weaver TL. DNA replication control through interaction of E2F-RB and the origin recognition complex. Nat Cell Biol. 2001;3:289–95. doi: 10.1038/35060086. [DOI] [PubMed] [Google Scholar]
- Bose ME, Mcconnell KH, Gardner-Aukema KA, Muller U, Weinreich M, Keck JL, Fox CA. The origin recognition complex and Sir4 protein recruit Sir1p to yeast silent chromatin through independent interactions requiring a common Sir1p domain. Mol Cell Biol. 2004;24:774–86. doi: 10.1128/MCB.24.2.774-786.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boskovic J, Bragado-Nilsson E, Saligram Prabhakar B, Yefimenko I, Martinez-Gago J, Munoz S, Mendez J, Montoya G. Molecular architecture of the recombinant human MCM2–7 helicase in complex with nucleotides and DNA. Cell Cycle. 2016:0. doi: 10.1080/15384101.2016.1191712. [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;8:678–84. doi: 10.1038/sj.embor.7401002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bowers JL, Randell JC, Chen S, Bell SP. ATP hydrolysis by ORC catalyzes reiterative Mcm2–7 assembly at a defined origin of replication. Mol Cell. 2004;16:967–78. doi: 10.1016/j.molcel.2004.11.038. [DOI] [PubMed] [Google Scholar]
- Brewer BJ, Fangman WL. The localization of replication origins on ARS plasmids in S. cerevisiae. Cell. 1987;51:463–71. doi: 10.1016/0092-8674(87)90642-8. [DOI] [PubMed] [Google Scholar]
- Brewster AS, Slaymaker IM, Afif SA, Chen XS. Mutational analysis of an archaeal minichromosome maintenance protein exterior hairpin reveals critical residues for helicase activity and DNA binding. BMC Mol Biol. 2010;11:62. doi: 10.1186/1471-2199-11-62. [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;105:20191–6. doi: 10.1073/pnas.0808037105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Broek D, Bartlett R, Crawford K, Nurse P. Involvement of p34cdc2 in establishing the dependency of S phase on mitosis. Nature. 1991;349:388–93. doi: 10.1038/349388a0. [DOI] [PubMed] [Google Scholar]
- Bruck I, Kanter DM, Kaplan DL. Enabling association of the GINS protein tetramer with the mini chromosome maintenance (Mcm)2–7 protein complex by phosphorylated Sld2 protein and single-stranded origin DNA. J Biol Chem. 2011;286:36414–26. doi: 10.1074/jbc.M111.282822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bruck I, Kaplan D. Dbf4-Cdc7 phosphorylation of Mcm2 is required for cell growth. J Biol Chem. 2009;284:28823–31. doi: 10.1074/jbc.M109.039123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bruck I, Kaplan DL. Origin single-stranded DNA releases Sld3 protein from the Mcm2–7 complex, allowing the GINS tetramer to bind the Mcm2–7 complex. J Biol Chem. 2011;286:18602–13. doi: 10.1074/jbc.M111.226332. [DOI] [PMC free article] [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;288:7550–63. doi: 10.1074/jbc.M112.440941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bruck I, Kaplan DL. The replication initiation protein Sld2 regulates helicase assembly. J Biol Chem. 2014;289:1948–59. doi: 10.1074/jbc.M113.532085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bruck I, Kaplan DL. Conserved mechanism for coordinating replication fork helicase assembly with phosphorylation of the helicase. Proc Natl Acad Sci U S A. 2015a;112:11223–8. doi: 10.1073/pnas.1509608112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bruck I, Kaplan DL. The Dbf4-Cdc7 kinase promotes Mcm2–7 ring opening to allow for single-stranded DNA extrusion and helicase assembly. J Biol Chem. 2015b;290:1210–21. doi: 10.1074/jbc.M114.608232. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- Bruck I, Kaplan DL. The Replication Initiation Protein Sld3/Treslin Orchestrates the Assembly of the Replication Fork Helicase during S Phase. J Biol Chem. 2015c;290:27414–24. doi: 10.1074/jbc.M115.688424. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- Buchman AR, Kimmerly WJ, Rine J, Kornberg RD. Two DNA-binding factors recognize specific sequences at silencers, upstream activating sequences, autonomously replicating sequences, and telomeres in Saccharomyces cerevisiae. Mol Cell Biol. 1988;8:210–25. doi: 10.1128/mcb.8.1.210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burkhart R, Schulte D, Hu D, Musahl C, Gohring F, Knippers R. Interactions of human nuclear proteins P1Mcm3 and P1Cdc46. Eur J Biochem. 1995;228:431–8. [PubMed] [Google Scholar]
- Cadoret JC, Meisch F, Hassan-Zadeh V, Luyten I, Guillet C, Duret L, Quesneville H, Prioleau MN. Genome-wide studies highlight indirect links between human replication origins and gene regulation. Proc Natl Acad Sci U S A. 2008;105:15837–42. doi: 10.1073/pnas.0805208105. [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;19:1905–19. doi: 10.1101/gad.337205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carpenter PB, Mueller PR, Dunphy WG. Role for a Xenopus Orc2-related protein in controlling DNA replication. Nature. 1996;379:357–60. doi: 10.1038/379357a0. [DOI] [PubMed] [Google Scholar]
- Cayrou C, Ballester B, Peiffer I, Fenouil R, Coulombe P, Andrau JC, Van Helden J, Mechali M. The chromatin environment shapes DNA replication origin organization and defines origin classes. Genome Res. 2015;25:1873–85. doi: 10.1101/gr.192799.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cayrou C, Coulombe P, Puy A, Rialle S, Kaplan N, Segal E, Mechali M. New insights into replication origin characteristics in metazoans. Cell Cycle. 2012;11:658–67. doi: 10.4161/cc.11.4.19097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cayrou C, Coulombe P, Vigneron A, Stanojcic S, Ganier O, Peiffer I, Rivals E, Puy A, Laurent-Chabalier S, Desprat R, Mechali M. Genome-scale analysis of metazoan replication origins reveals their organization in specific but flexible sites defined by conserved features. Genome Res. 2011;21:1438–49. doi: 10.1101/gr.121830.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chakraborty A, Shen Z, Prasanth SG. “ORCanization” on heterochromatin: linking DNA replication initiation to chromatin organization. Epigenetics. 2011;6:665–70. doi: 10.4161/epi.6.6.16179. [DOI] [PubMed] [Google Scholar]
- Chang F, Riera A, Evrin C, Sun J, Li H, Speck C, Weinreich M. Cdc6 ATPase activity disengages Cdc6 from the pre-replicative complex to promote DNA replication. Elife. 2015:4. doi: 10.7554/eLife.05795. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chapman JW, Johnston LH. The yeast gene, DBF4, essential for entry into S phase is cell cycle regulated. Exp Cell Res. 1989;180:419–28. doi: 10.1016/0014-4827(89)90068-2. [DOI] [PubMed] [Google Scholar]
- Chen S, Bell SP. CDK prevents Mcm2–7 helicase loading by inhibiting Cdt1 interaction with Orc6. Genes Dev. 2011;25:363–72. doi: 10.1101/gad.2011511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen S, De Vries MA, Bell SP. Orc6 is required for dynamic recruitment of Cdt1 during repeated Mcm2–7 loading. Genes Dev. 2007;21:2897–907. doi: 10.1101/gad.1596807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen Y, Hennessy KM, Botstein D, Tye BK. CDC46/MCM5, a yeast protein whose subcellular localization is cell cycle-regulated, is involved in DNA replication at autonomously replicating sequences. Proc Natl Acad Sci U S A. 1992;89:10459–63. doi: 10.1073/pnas.89.21.10459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen YJ, Yu X, Kasiviswanathan R, Shin JH, Kelman Z, Egelman EH. Structural polymorphism of Methanothermobacter thermautotrophicus MCM. J Mol Biol. 2005;346:389–94. doi: 10.1016/j.jmb.2004.11.076. [DOI] [PubMed] [Google Scholar]
- Chen Z, Speck C, Wendel P, Tang C, Stillman B, Li H. The architecture of the DNA replication origin recognition complex in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2008;105:10326–31. doi: 10.1073/pnas.0803829105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng L, Collyer T, Hardy CF. Cell cycle regulation of DNA replication initiator factor Dbf4p. Mol Cell Biol. 1999;19:4270–8. doi: 10.1128/mcb.19.6.4270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chesnokov I, Gossen M, Remus D, Botchan M. Assembly of functionally active Drosophila origin recognition complex from recombinant proteins. Genes Dev. 1999;13:1289–96. doi: 10.1101/gad.13.10.1289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chesnokov I, Remus D, Botchan M. Functional analysis of mutant and wild-type Drosophila origin recognition complex. Proc Natl Acad Sci U S A. 2001;98:11997–2002. doi: 10.1073/pnas.211342798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chesnokov IN. Multiple functions of the origin recognition complex. Int Rev Cytol. 2007;256:69–109. doi: 10.1016/S0074-7696(07)56003-1. [DOI] [PubMed] [Google Scholar]
- Chesnokov IN, Chesnokova ON, Botchan M. A cytokinetic function of Drosophila ORC6 protein resides in a domain distinct from its replication activity. Proc Natl Acad Sci U S A. 2003;100:9150–5. doi: 10.1073/pnas.1633580100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho WH, Lee YJ, Kong SI, Hurwitz J, Lee JK. CDC7 kinase phosphorylates serine residues adjacent to acidic amino acids in the minichromosome maintenance 2 protein. Proc Natl Acad Sci U S A. 2006;103:11521–6. doi: 10.1073/pnas.0604990103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chong JP, Hayashi MK, Simon MN, Xu RM, Stillman B. A double-hexamer archaeal minichromosome maintenance protein is an ATP-dependent DNA helicase. Proc Natl Acad Sci U S A. 2000;97:1530–5. doi: 10.1073/pnas.030539597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chong JP, Mahbubani HM, Khoo CY, Blow JJ. Purification of an MCM-containing complex as a component of the DNA replication licensing system. Nature. 1995;375:418–21. doi: 10.1038/375418a0. [DOI] [PubMed] [Google Scholar]
- Chong JP, Thommes P, Blow JJ. The role of MCM/P1 proteins in the licensing of DNA replication. Trends Biochem Sci. 1996;21:102–6. [PubMed] [Google Scholar]
- Chuang RY, Chretien L, Dai J, Kelly TJ. Purification and characterization of the Schizosaccharomyces pombe origin recognition complex: interaction with origin DNA and Cdc18 protein. J Biol Chem. 2002;277:16920–7. doi: 10.1074/jbc.M107710200. [DOI] [PubMed] [Google Scholar]
- Chuang RY, Kelly TJ. The fission yeast homologue of Orc4p binds to replication origin DNA via multiple AT-hooks. Proc Natl Acad Sci U S A. 1999;96:2656–61. doi: 10.1073/pnas.96.6.2656. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clarey MG, Botchan M, Nogales E. Single particle EM studies of the Drosophila melanogaster origin recognition complex and evidence for DNA wrapping. J Struct Biol. 2008;164:241–9. doi: 10.1016/j.jsb.2008.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clarey MG, Erzberger JP, Grob P, Leschziner AE, Berger JM, Nogales E, Botchan M. Nucleotide-dependent conformational changes in the DnaA-like core of the origin recognition complex. Nat Struct Mol Biol. 2006;13:684–90. doi: 10.1038/nsmb1121. [DOI] [PubMed] [Google Scholar]
- Claycomb JM, Macalpine DM, Evans JG, Bell SP, Orr-Weaver TL. Visualization of replication initiation and elongation in Drosophila. J Cell Biol. 2002;159:225–36. doi: 10.1083/jcb.200207046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clyne RK, Kelly TJ. Genetic analysis of an ARS element from the fission yeast Schizosaccharomyces pombe. EMBO J. 1995;14:6348–57. doi: 10.1002/j.1460-2075.1995.tb00326.x. [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;379:180–2. doi: 10.1038/379180a0. [DOI] [PubMed] [Google Scholar]
- Coleman TR, Carpenter PB, Dunphy WG. The Xenopus Cdc6 protein is essential for the initiation of a single round of DNA replication in cell-free extracts. Cell. 1996;87:53–63. doi: 10.1016/s0092-8674(00)81322-7. [DOI] [PubMed] [Google Scholar]
- Cook JG, Chasse DA, Nevins JR. The regulated association of Cdt1 with minichromosome maintenance proteins and Cdc6 in mammalian cells. J Biol Chem. 2004;279:9625–33. doi: 10.1074/jbc.M311933200. [DOI] [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, 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;18:471–7. doi: 10.1038/nsmb.2004. [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]
- Costa A, Van Duinen G, Medagli B, Chong J, Sakakibara N, Kelman Z, Nair SK, Patwardhan A, Onesti S. Cryo-electron microscopy reveals a novel DNA-binding site on the MCM helicase. EMBO J. 2008;27:2250–8. doi: 10.1038/emboj.2008.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costas C, De La Paz Sanchez M, Stroud H, Yu Y, Oliveros JC, Feng S, Benguria A, Lopez-Vidriero I, Zhang X, Solano R, Jacobsen SE, Gutierrez C. Genome-wide mapping of Arabidopsis thaliana origins of DNA replication and their associated epigenetic marks. Nat Struct Mol Biol. 2011;18:395–400. doi: 10.1038/nsmb.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coster G, Frigola J, Beuron F, Morris EP, Diffley JF. Origin licensing requires ATP binding and hydrolysis by the MCM replicative helicase. Mol Cell. 2014;55:666–77. doi: 10.1016/j.molcel.2014.06.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Creager RL, Li Y, Macalpine DM. SnapShot: Origins of DNA replication. Cell. 2015;161:418–418. e1. doi: 10.1016/j.cell.2015.03.043. [DOI] [PubMed] [Google Scholar]
- Crevel G, Ivetic A, Ohno K, Yamaguchi M, Cotterill S. Nearest neighbour analysis of MCM protein complexes in Drosophila melanogaster. Nucleic Acids Res. 2001;29:4834–42. doi: 10.1093/nar/29.23.4834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dai J, Chuang RY, Kelly TJ. DNA replication origins in the Schizosaccharomyces pombe genome. Proc Natl Acad Sci U S A. 2005;102:337–42. doi: 10.1073/pnas.0408811102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalton S, Hopwood B. Characterization of Cdc47p-minichromosome maintenance complexes in Saccharomyces cerevisiae: identification of Cdc45p as a subunit. Mol Cell Biol. 1997;17:5867–75. doi: 10.1128/mcb.17.10.5867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalton S, Whitbread L. Cell cycle-regulated nuclear import and export of Cdc47, a protein essential for initiation of DNA replication in budding yeast. Proc Natl Acad Sci U S A. 1995;92:2514–8. doi: 10.1073/pnas.92.7.2514. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davey MJ, Indiani C, O’donnell M. Reconstitution of the Mcm2–7p heterohexamer, subunit arrangement, and ATP site architecture. J Biol Chem. 2003;278:4491–9. doi: 10.1074/jbc.M210511200. [DOI] [PubMed] [Google Scholar]
- De Marco V, Gillespie PJ, Li A, Karantzelis N, Christodoulou E, Klompmaker R, Van Gerwen S, Fish A, Petoukhov MV, Iliou MS, Lygerou Z, Medema RH, Blow JJ, Svergun DI, Taraviras S, Perrakis A. Quaternary structure of the human Cdt1-Geminin complex regulates DNA replication licensing. Proc Natl Acad Sci U S A. 2009;106:19807–12. doi: 10.1073/pnas.0905281106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deegan TD, Yeeles JT, Diffley JF. Phosphopeptide binding by Sld3 links Dbf4-dependent kinase to MCM replicative helicase activation. EMBO J. 2016;35:961–73. doi: 10.15252/embj.201593552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Delgado S, Gomez M, Bird A, Antequera F. Initiation of DNA replication at CpG islands in mammalian chromosomes. EMBO J. 1998;17:2426–35. doi: 10.1093/emboj/17.8.2426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng Z, Norseen J, Wiedmer A, Riethman H, Lieberman PM. TERRA RNA binding to TRF2 facilitates heterochromatin formation and ORC recruitment at telomeres. Mol Cell. 2009;35:403–13. doi: 10.1016/j.molcel.2009.06.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devault A, Gueydon E, Schwob E. Interplay between S-cyclin-dependent kinase and Dbf4-dependent kinase in controlling DNA replication through phosphorylation of yeast Mcm4 N-terminal domain. Mol Biol Cell. 2008;19:2267–77. doi: 10.1091/mbc.E07-06-0614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devault A, Vallen EA, Yuan T, Green S, Bensimon A, Schwob E. Identification of Tah11/Sid2 as the ortholog of the replication licensing factor Cdt1 in Saccharomyces cerevisiae. Curr Biol. 2002;12:689–94. doi: 10.1016/s0960-9822(02)00768-6. [DOI] [PubMed] [Google Scholar]
- Dewar JM, Budzowska M, Walter JC. The mechanism of DNA replication termination in vertebrates. Nature. 2015;525:345–50. doi: 10.1038/nature14887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dhar SK, Delmolino L, Dutta A. Architecture of the human origin recognition complex. J Biol Chem. 2001;276:29067–71. doi: 10.1074/jbc.M103078200. [DOI] [PubMed] [Google Scholar]
- Dhar SK, Dutta A. Identification and characterization of the human ORC6 homolog. J Biol Chem. 2000;275:34983–8. doi: 10.1074/jbc.M006069200. [DOI] [PubMed] [Google Scholar]
- Dhingra N, Bruck I, Smith S, Ning B, Kaplan DL. Dpb11 protein helps control assembly of the Cdc45.Mcm2–7.GINS replication fork helicase. J Biol Chem. 2015;290:7586–601. doi: 10.1074/jbc.M115.640383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diffley JF, Cocker JH. Protein-DNA interactions at a yeast replication origin. Nature. 1992;357:169–72. doi: 10.1038/357169a0. [DOI] [PubMed] [Google Scholar]
- Diffley JF, Cocker JH, Dowell SJ, Rowley A. Two steps in the assembly of complexes at yeast replication origins in vivo. Cell. 1994;78:303–16. doi: 10.1016/0092-8674(94)90299-2. [DOI] [PubMed] [Google Scholar]
- Diffley JF, Stillman B. Purification of a yeast protein that binds to origins of DNA replication and a transcriptional silencer. Proc Natl Acad Sci U S A. 1988;85:2120–4. doi: 10.1073/pnas.85.7.2120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dimaki M, Xouri G, Symeonidou IE, Sirinian C, Nishitani H, Taraviras S, Lygerou Z. Cell cycle-dependent subcellular translocation of the human DNA licensing inhibitor geminin. J Biol Chem. 2013;288:23953–63. doi: 10.1074/jbc.M113.453092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Donovan JD, Toyn JH, Johnson AL, Johnston LH. P40SDB25, a putative CDK inhibitor, has a role in the M/G1 transition in Saccharomyces cerevisiae. Genes Dev. 1994;8:1640–53. doi: 10.1101/gad.8.14.1640. [DOI] [PubMed] [Google Scholar]
- Donovan S, Harwood J, Drury LS, Diffley JF. Cdc6p-dependent loading of Mcm proteins onto pre-replicative chromatin in budding yeast. Proc Natl Acad Sci U S A. 1997;94:5611–6. doi: 10.1073/pnas.94.11.5611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Douglas ME, Diffley JF. Recruitment of Mcm10 to Sites of Replication Initiation Requires Direct Binding to the Minichromosome Maintenance (MCM) Complex. J Biol Chem. 2016;291:5879–88. doi: 10.1074/jbc.M115.707802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dowell SJ, Romanowski P, Diffley JF. Interaction of Dbf4, the Cdc7 protein kinase regulatory subunit, with yeast replication origins in vivo. Science. 1994;265:1243–6. doi: 10.1126/science.8066465. [DOI] [PubMed] [Google Scholar]
- Drury LS, Perkins G, Diffley JF. The Cdc4/34/53 pathway targets Cdc6p for proteolysis in budding yeast. EMBO J. 1997;16:5966–76. doi: 10.1093/emboj/16.19.5966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dueber EC, Costa A, Corn JE, Bell SD, Berger JM. Molecular determinants of origin discrimination by Orc1 initiators in archaea. Nucleic Acids Res. 2011;39:3621–31. doi: 10.1093/nar/gkq1308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dueber EL, Corn JE, Bell SD, Berger JM. Replication origin recognition and deformation by a heterodimeric archaeal Orc1 complex. Science. 2007;317:1210–3. doi: 10.1126/science.1143690. [DOI] [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;58:483–94. doi: 10.1016/j.molcel.2015.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eaton ML, Galani K, Kang S, Bell SP, Macalpine DM. Conserved nucleosome positioning defines replication origins. Genes Dev. 2010;24:748–53. doi: 10.1101/gad.1913210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eaton ML, Prinz JA, Macalpine HK, Tretyakov G, Kharchenko PV, Macalpine DM. Chromatin signatures of the Drosophila replication program. Genome Res. 2011;21:164–74. doi: 10.1101/gr.116038.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edgell DR, Doolittle WF. Archaea and the origin(s) of DNA replication proteins. Cell. 1997;89:995–8. doi: 10.1016/s0092-8674(00)80285-8. [DOI] [PubMed] [Google Scholar]
- Edwards MC, Tutter AV, Cvetic C, Gilbert CH, Prokhorova TA, Walter JC. MCM2–7 complexes bind chromatin in a distributed pattern surrounding the origin recognition complex in Xenopus egg extracts. J Biol Chem. 2002;277:33049–57. doi: 10.1074/jbc.M204438200. [DOI] [PubMed] [Google Scholar]
- El-Sayed NM, Myler PJ, Bartholomeu DC, Nilsson D, Aggarwal G, Tran AN, Ghedin E, Worthey EA, Delcher AL, Blandin G, Westenberger SJ, Caler E, Cerqueira GC, Branche C, Haas B, Anupama A, Arner E, Aslund L, Attipoe P, Bontempi E, Bringaud F, Burton P, Cadag E, Campbell DA, Carrington M, Crabtree J, Darban H, Da Silveira JF, De Jong P, Edwards K, Englund PT, Fazelina G, Feldblyum T, Ferella M, Frasch AC, Gull K, Horn D, Hou L, Huang Y, Kindlund E, Klingbeil M, Kluge S, Koo H, Lacerda D, Levin MJ, Lorenzi H, Louie T, Machado CR, Mcculloch R, Mckenna A, Mizuno Y, Mottram JC, Nelson S, Ochaya S, Osoegawa K, Pai G, Parsons M, Pentony M, Pettersson U, Pop M, Ramirez JL, Rinta J, Robertson L, Salzberg SL, Sanchez DO, Seyler A, Sharma R, Shetty J, Simpson AJ, Sisk E, Tammi MT, Tarleton R, Teixeira S, Van Aken S, Vogt C, Ward PN, Wickstead B, Wortman J, White O, Fraser CM, Stuart KD, Andersson B. The genome sequence of Trypanosoma cruzi, etiologic agent of Chagas disease. Science. 2005;309:409–15. doi: 10.1126/science.1112631. [DOI] [PubMed] [Google Scholar]
- Enemark EJ, Joshua-Tor L. Mechanism of DNA translocation in a replicative hexameric helicase. Nature. 2006;442:270–5. doi: 10.1038/nature04943. [DOI] [PubMed] [Google Scholar]
- Enemark EJ, Joshua-Tor L. On helicases and other motor proteins. Curr Opin Struct Biol. 2008;18:243–57. doi: 10.1016/j.sbi.2008.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erzberger JP, Berger JM. Evolutionary relationships and structural mechanisms of AAA+ proteins. Annu Rev Biophys Biomol Struct. 2006;35:93–114. doi: 10.1146/annurev.biophys.35.040405.101933. [DOI] [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;106:20240–5. doi: 10.1073/pnas.0911500106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evrin C, Fernandez-Cid A, Riera A, Zech J, Clarke P, Herrera MC, Tognetti S, Lurz R, Speck C. The ORC/Cdc6/MCM2–7 complex facilitates MCM2–7 dimerization during prereplicative complex formation. Nucleic Acids Res. 2014;42:2257–69. doi: 10.1093/nar/gkt1148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evrin C, Fernandez-Cid A, Zech J, Herrera MC, Riera A, Clarke P, Brill S, Lurz R, Speck C. In the absence of ATPase activity, pre-RC formation is blocked prior to MCM2–7 hexamer dimerization. Nucleic Acids Res. 2013;41:3162–72. doi: 10.1093/nar/gkt043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fang D, Cao Q, Lou H. Sld3-MCM Interaction Facilitated by Dbf4-Dependent Kinase Defines an Essential Step in Eukaryotic DNA Replication Initiation. Front Microbiol. 2016;7:885. doi: 10.3389/fmicb.2016.00885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng L, Hu Y, Wang B, Wu L, Jong A. Loss control of Mcm5 interaction with chromatin in cdc6–1 mutated in CDC-NTP motif. DNA Cell Biol. 2000;19:447–57. doi: 10.1089/10445490050085933. [DOI] [PubMed] [Google Scholar]
- Fenouil R, Cauchy P, Koch F, Descostes N, Cabeza JZ, Innocenti C, Ferrier P, Spicuglia S, Gut M, Gut I, Andrau JC. CpG islands and GC content dictate nucleosome depletion in a transcription-independent manner at mammalian promoters. Genome Res. 2012;22:2399–408. doi: 10.1101/gr.138776.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferenbach A, Li A, Brito-Martins M, Blow JJ. Functional domains of the Xenopus replication licensing factor Cdt1. Nucleic Acids Res. 2005;33:316–24. doi: 10.1093/nar/gki176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fernandez-Cid A, Riera A, Tognetti S, Herrera MC, Samel S, Evrin C, Winkler C, Gardenal E, Uhle S, Speck C. An ORC/Cdc6/MCM2–7 complex is formed in a multistep reaction to serve as a platform for MCM double-hexamer assembly. Mol Cell. 2013;50:577–88. doi: 10.1016/j.molcel.2013.03.026. [DOI] [PubMed] [Google Scholar]
- Findeisen M, El-Denary M, Kapitza T, Graf R, Strausfeld U. Cyclin A-dependent kinase activity affects chromatin binding of ORC, Cdc6, and MCM in egg extracts of Xenopus laevis. Eur J Biochem. 1999;264:415–26. doi: 10.1046/j.1432-1327.1999.00613.x. [DOI] [PubMed] [Google Scholar]
- Fiorani P, Bjornsti MA. Mechanisms of DNA topoisomerase I-induced cell killing in the yeast Saccharomyces cerevisiae. Ann N Y Acad Sci. 2000;922:65–75. doi: 10.1111/j.1749-6632.2000.tb07026.x. [DOI] [PubMed] [Google Scholar]
- Fletcher RJ, Bishop BE, Leon RP, Sclafani RA, Ogata CM, Chen XS. The structure and function of MCM from archaeal M. Thermoautotrophicum. Nat Struct Biol. 2003;10:160–7. doi: 10.1038/nsb893. [DOI] [PubMed] [Google Scholar]
- Fletcher RJ, Shen J, Gomez-Llorente Y, Martin CS, Carazo JM, Chen XS. Double hexamer disruption and biochemical activities of Methanobacterium thermoautotrophicum MCM. J Biol Chem. 2005;280:42405–10. doi: 10.1074/jbc.M509773200. [DOI] [PubMed] [Google Scholar]
- Fletcher RJ, Shen J, Holden LG, Chen XS. Identification of amino acids important for the biochemical activity of Methanothermobacter thermautotrophicus MCM. Biochemistry. 2008;47:9981–6. doi: 10.1021/bi800032t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Foss M, Mcnally FJ, Laurenson P, Rine J. Origin recognition complex (ORC) in transcriptional silencing and DNA replication in S. cerevisiae. Science. 1993;262:1838–44. doi: 10.1126/science.8266071. [DOI] [PubMed] [Google Scholar]
- Fox CA, Ehrenhofer-Murray AE, Loo S, Rine J. The origin recognition complex, SIR1, and the S phase requirement for silencing. Science. 1997;276:1547–51. doi: 10.1126/science.276.5318.1547. [DOI] [PubMed] [Google Scholar]
- Fox CA, Loo S, Dillin A, Rine J. The origin recognition complex has essential functions in transcriptional silencing and chromosomal replication. Genes Dev. 1995;9:911–24. doi: 10.1101/gad.9.8.911. [DOI] [PubMed] [Google Scholar]
- Francis LI, Randell JC, Takara TJ, Uchima L, Bell SP. Incorporation into the prereplicative complex activates the Mcm2–7 helicase for Cdc7-Dbf4 phosphorylation. Genes Dev. 2009;23:643–54. doi: 10.1101/gad.1759609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Francois Jacob SB, Francois Cuzin. On the Regulation of DNA Replication in Bacteria. Cold Spring Harb Symp Quant Biol. 1963;28:329–348. [Google Scholar]
- Frigola J, Remus D, Mehanna A, Diffley JF. ATPase-dependent quality control of DNA replication origin licensing. Nature. 2013;495:339–43. doi: 10.1038/nature11920. [DOI] [PMC free article] [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 Y, Slaymaker IM, Wang J, Wang G, Chen XS. The 1.8-A crystal structure of the N-terminal domain of an archaeal MCM as a right-handed filament. J Mol Biol. 2014;426:1512–23. doi: 10.1016/j.jmb.2013.12.025. [DOI] [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;146:931–41. doi: 10.1016/j.cell.2011.07.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujita M, Hori Y, Shirahige K, Tsurimoto T, Yoshikawa H, Obuse C. Cell cycle dependent topological changes of chromosomal replication origins in Saccharomyces cerevisiae. Genes Cells. 1998;3:737–49. doi: 10.1046/j.1365-2443.1998.00226.x. [DOI] [PubMed] [Google Scholar]
- Gaczynska M, Osmulski PA, Jiang Y, Lee JK, Bermudez V, Hurwitz J. Atomic force microscopic analysis of the binding of the Schizosaccharomyces pombe origin recognition complex and the spOrc4 protein with origin DNA. Proc Natl Acad Sci U S A. 2004;101:17952–7. doi: 10.1073/pnas.0408369102. [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;8:358–66. doi: 10.1038/ncb1382. [DOI] [PubMed] [Google Scholar]
- Gambus A, Khoudoli GA, Jones RC, Blow JJ. MCM2–7 form double hexamers at licensed origins in Xenopus egg extract. J Biol Chem. 2011;286:11855–64. doi: 10.1074/jbc.M110.199521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gardner KA, Rine J, Fox CA. A region of the Sir1 protein dedicated to recognition of a silencer and required for interaction with the Orc1 protein in saccharomyces cerevisiae. Genetics. 1999;151:31–44. doi: 10.1093/genetics/151.1.31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gaudier M, Schuwirth BS, Westcott SL, Wigley DB. Structural basis of DNA replication origin recognition by an ORC protein. Science. 2007;317:1213–6. doi: 10.1126/science.1143664. [DOI] [PubMed] [Google Scholar]
- Gavin KA, Hidaka M, Stillman B. Conserved initiator proteins in eukaryotes. Science. 1995;270:1667–71. doi: 10.1126/science.270.5242.1667. [DOI] [PubMed] [Google Scholar]
- Gibson SI, Surosky RT, Tye BK. The phenotype of the minichromosome maintenance mutant mcm3 is characteristic of mutants defective in DNA replication. Mol Cell Biol. 1990;10:5707–20. doi: 10.1128/mcb.10.11.5707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilbert DM. In search of the holy replicator. Nat Rev Mol Cell Biol. 2004;5:848–55. doi: 10.1038/nrm1495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gillespie PJ, Li A, Blow JJ. Reconstitution of licensed replication origins on Xenopus sperm nuclei using purified proteins. BMC Biochem. 2001;2:15. doi: 10.1186/1471-2091-2-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giordano-Coltart J, Ying CY, Gautier J, Hurwitz J. Studies of the properties of human origin recognition complex and its Walker A motif mutants. Proc Natl Acad Sci U S A. 2005;102:69–74. doi: 10.1073/pnas.0408690102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Givens RM, Lai WK, Rizzo JM, Bard JE, Mieczkowski PA, Leatherwood J, Huberman JA, Buck MJ. Chromatin architectures at fission yeast transcriptional promoters and replication origins. Nucleic Acids Res. 2012;40:7176–89. doi: 10.1093/nar/gks351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gomez-Llorente Y, Fletcher RJ, Chen XS, Carazo JM, San Martin C. Polymorphism and double hexamer structure in the archaeal minichromosome maintenance (MCM) helicase from Methanobacterium thermoautotrophicum. J Biol Chem. 2005;280:40909–15. doi: 10.1074/jbc.M509760200. [DOI] [PubMed] [Google Scholar]
- Gopalakrishnan V, Simancek P, Houchens C, Snaith HA, Frattini MG, Sazer S, Kelly TJ. Redundant control of rereplication in fission yeast. Proc Natl Acad Sci U S A. 2001;98:13114–9. doi: 10.1073/pnas.221467598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gossen M, Pak DT, Hansen SK, Acharya JK, Botchan MR. A Drosophila homolog of the yeast origin recognition complex. Science. 1995;270:1674–7. doi: 10.1126/science.270.5242.1674. [DOI] [PubMed] [Google Scholar]
- Graham BW, Schauer GD, Leuba SH, Trakselis MA. Steric exclusion and wrapping of the excluded DNA strand occurs along discrete external binding paths during MCM helicase unwinding. Nucleic Acids Res. 2011;39:6585–95. doi: 10.1093/nar/gkr345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Graham BW, Tao Y, Dodge KL, Thaxton CT, Olaso D, Young NL, Marshall AG, Trakselis MA. DNA Interactions Probed by H/D Exchange FT-ICR Mass Spectrometry Confirm External Binding Sites on the MCM Helicase. J Biol Chem. 2016 doi: 10.1074/jbc.M116.719591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grallert B, Nurse P. The ORC1 homolog orp1 in fission yeast plays a key role in regulating onset of S phase. Genes Dev. 1996;10:2644–54. doi: 10.1101/gad.10.20.2644. [DOI] [PubMed] [Google Scholar]
- Gregan J, Lindner K, Brimage L, Franklin R, Namdar M, Hart EA, Aves SJ, Kearsey SE. Fission yeast Cdc23/Mcm10 functions after pre-replicative complex formation to promote Cdc45 chromatin binding. Mol Biol Cell. 2003;14:3876–87. doi: 10.1091/mbc.E03-02-0090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gros J, Kumar C, Lynch G, Yadav T, Whitehouse I, Remus D. Post-licensing Specification of Eukaryotic Replication Origins by Facilitated Mcm2–7 Sliding along DNA. Molecular Cell. 2015;60:797–807. doi: 10.1016/j.molcel.2015.10.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guarino E, Shepherd ME, Salguero I, Hua H, Deegan RS, Kearsey SE. Cdt1 proteolysis is promoted by dual PIP degrons and is modulated by PCNA ubiquitylation. Nucleic Acids Res. 2011;39:5978–90. doi: 10.1093/nar/gkr222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guenther B, Onrust R, Sali A, O’donnell M, Kuriyan J. Crystal structure of the delta’ subunit of the clamp-loader complex of E. coli DNA polymerase III. Cell. 1997;91:335–45. doi: 10.1016/s0092-8674(00)80417-1. [DOI] [PubMed] [Google Scholar]
- Hanson PI, Whiteheart SW. AAA+ proteins: have engine, will work. Nat Rev Mol Cell Biol. 2005;6:519–29. doi: 10.1038/nrm1684. [DOI] [PubMed] [Google Scholar]
- Hardy CF, Dryga O, Seematter S, Pahl PM, Sclafani RA. mcm5/cdc46-bob1 bypasses the requirement for the S phase activator Cdc7p. Proc Natl Acad Sci U S A. 1997;94:3151–5. doi: 10.1073/pnas.94.7.3151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hartwell LH. Sequential function of gene products relative to DNA synthesis in the yeast cell cycle. J Mol Biol. 1976;104:803–17. doi: 10.1016/0022-2836(76)90183-2. [DOI] [PubMed] [Google Scholar]
- Harvey KJ, Newport J. Metazoan origin selection: origin recognition complex chromatin binding is regulated by CDC6 recruitment and ATP hydrolysis. J Biol Chem. 2003;278:48524–8. doi: 10.1074/jbc.M307661200. [DOI] [PubMed] [Google Scholar]
- Havens CG, Walter JC. Docking of a specialized PIP Box onto chromatin-bound PCNA creates a degron for the ubiquitin ligase CRL4Cdt2. Mol Cell. 2009;35:93–104. doi: 10.1016/j.molcel.2009.05.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Havens CG, Walter JC. Mechanism of CRL4(Cdt2), a PCNA-dependent E3 ubiquitin ligase. Genes Dev. 2011;25:1568–82. doi: 10.1101/gad.2068611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hayles J, Fisher D, Woollard A, Nurse P. Temporal order of S phase and mitosis in fission yeast is determined by the state of the p34cdc2-mitotic B cyclin complex. Cell. 1994;78:813–22. doi: 10.1016/s0092-8674(94)90542-8. [DOI] [PubMed] [Google Scholar]
- Heichinger C, Penkett CJ, Bahler J, Nurse P. Genome-wide characterization of fission yeast DNA replication origins. EMBO J. 2006;25:5171–9. doi: 10.1038/sj.emboj.7601390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heinzel SS, Krysan PJ, Tran CT, Calos MP. Autonomous DNA replication in human cells is affected by the size and the source of the DNA. Mol Cell Biol. 1991;11:2263–72. doi: 10.1128/mcb.11.4.2263. [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]
- Hennessy KM, Clark CD, Botstein D. Subcellular localization of yeast CDC46 varies with the cell cycle. Genes Dev. 1990;4:2252–63. doi: 10.1101/gad.4.12b.2252. [DOI] [PubMed] [Google Scholar]
- Hernandez-Perez S, Cabrera E, Amoedo H, Rodriguez-Acebes S, Koundrioukoff S, Debatisse M, Mendez J, Freire R. USP37 deubiquitinates Cdt1 and contributes to regulate DNA replication. Mol Oncol. 2016;10:1196–206. doi: 10.1016/j.molonc.2016.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Higa M, Kushiyama T, Kurashige S, Kohmon D, Enokitani K, Iwahori S, Sugimoto N, Yoshida K, Fujita M. TRF2 recruits ORC through TRFH domain dimerization. Biochim Biophys Acta. 2016;1864:191–201. doi: 10.1016/j.bbamcr.2016.11.004. [DOI] [PubMed] [Google Scholar]
- Hizume K, Yagura M, Araki H. Concerted interaction between origin recognition complex (ORC), nucleosomes and replication origin DNA ensures stable ORC-origin binding. Genes Cells. 2013;18:764–79. doi: 10.1111/gtc.12073. [DOI] [PubMed] [Google Scholar]
- Hoang ML, Leon RP, Pessoa-Brandao L, Hunt S, Raghuraman MK, Fangman WL, Brewer BJ, Sclafani RA. Structural changes in Mcm5 protein bypass Cdc7-Dbf4 function and reduce replication origin efficiency in Saccharomyces cerevisiae. Mol Cell Biol. 2007;27:7594–602. doi: 10.1128/MCB.00997-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hofmann JF, Beach D. cdt1 is an essential target of the Cdc10/Sct1 transcription factor: requirement for DNA replication and inhibition of mitosis. EMBO J. 1994;13:425–34. doi: 10.1002/j.1460-2075.1994.tb06277.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hogan E, Koshland D. Addition of extra origins of replication to a minichromosome suppresses its mitotic loss in cdc6 and cdc14 mutants of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 1992;89:3098–102. doi: 10.1073/pnas.89.7.3098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hoggard T, Shor E, Muller CA, Nieduszynski CA, Fox CA. A Link between ORC-origin binding mechanisms and origin activation time revealed in budding yeast. PLoS Genet. 2013;9:e1003798. doi: 10.1371/journal.pgen.1003798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hollingsworth RE, Jr, Sclafani RA. DNA metabolism gene CDC7 from yeast encodes a serine (threonine) protein kinase. Proc Natl Acad Sci U S A. 1990;87:6272–6. doi: 10.1073/pnas.87.16.6272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Homesley L, Lei M, Kawasaki Y, Sawyer S, Christensen T, Tye BK. Mcm10 and the MCM2–7 complex interact to initiate DNA synthesis and to release replication factors from origins. Genes Dev. 2000;14:913–26. [PMC free article] [PubMed] [Google Scholar]
- Honey S, Futcher B. Roles of the CDK phosphorylation sites of yeast Cdc6 in chromatin binding and rereplication. Mol Biol Cell. 2007;18:1324–36. doi: 10.1091/mbc.E06-06-0544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hook SS, Lin JJ, Dutta A. Mechanisms to control rereplication and implications for cancer. Curr Opin Cell Biol. 2007;19:663–71. doi: 10.1016/j.ceb.2007.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hopwood B, Dalton S. Cdc45p assembles into a complex with Cdc46p/Mcm5p, is required for minichromosome maintenance, and is essential for chromosomal DNA replication. Proc Natl Acad Sci U S A. 1996;93:12309–14. doi: 10.1073/pnas.93.22.12309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hoshina S, Yura K, Teranishi H, Kiyasu N, Tominaga A, Kadoma H, Nakatsuka A, Kunichika T, Obuse C, Waga S. Human origin recognition complex binds preferentially to G-quadruplex-preferable RNA and single-stranded DNA. J Biol Chem. 2013;288:30161–71. doi: 10.1074/jbc.M113.492504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Houchens CR, Lu W, Chuang RY, Frattini MG, Fuller A, Simancek P, Kelly TJ. Multiple mechanisms contribute to Schizosaccharomyces pombe origin recognition complex-DNA interactions. J Biol Chem. 2008;283:30216–24. doi: 10.1074/jbc.M802649200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J, Xiong Y. An evolutionarily conserved function of proliferating cell nuclear antigen for Cdt1 degradation by the Cul4-Ddb1 ubiquitin ligase in response to DNA damage. J Biol Chem. 2006;281:3753–6. doi: 10.1074/jbc.C500464200. [DOI] [PubMed] [Google Scholar]
- Hua XH, Newport J. Identification of a preinitiation step in DNA replication that is independent of origin recognition complex and cdc6, but dependent on cdk2. J Cell Biol. 1998;140:271–81. doi: 10.1083/jcb.140.2.271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang DW, Fanti L, Pak DT, Botchan MR, Pimpinelli S, Kellum R. Distinct cytoplasmic and nuclear fractions of Drosophila heterochromatin protein 1: their phosphorylation levels and associations with origin recognition complex proteins. J Cell Biol. 1998;142:307–18. doi: 10.1083/jcb.142.2.307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huberman JA, Spotila LD, Nawotka KA, El-Assouli SM, Davis LR. The in vivo replication origin of the yeast 2 microns plasmid. Cell. 1987;51:473–81. doi: 10.1016/0092-8674(87)90643-x. [DOI] [PubMed] [Google Scholar]
- Huberman JA, Zhu JG, Davis LR, Newlon CS. Close association of a DNA replication origin and an ARS element on chromosome III of the yeast, Saccharomyces cerevisiae. Nucleic Acids Res. 1988;16:6373–84. doi: 10.1093/nar/16.14.6373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huppert JL, Balasubramanian S. G-quadruplexes in promoters throughout the human genome. Nucleic Acids Res. 2007;35:406–13. doi: 10.1093/nar/gkl1057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hyrien O, Mechali M. Chromosomal replication initiates and terminates at random sequences but at regular intervals in the ribosomal DNA of Xenopus early embryos. EMBO J. 1993;12:4511–20. doi: 10.1002/j.1460-2075.1993.tb06140.x. [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;37:247–58. doi: 10.1016/j.molcel.2009.12.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ishimi Y. A DNA helicase activity is associated with an MCM4, -6, and -7 protein complex. J Biol Chem. 1997;272:24508–13. doi: 10.1074/jbc.272.39.24508. [DOI] [PubMed] [Google Scholar]
- Iyer LM, Leipe DD, Koonin EV, Aravind L. Evolutionary history and higher order classification of AAA+ ATPases. J Struct Biol. 2004;146:11–31. doi: 10.1016/j.jsb.2003.10.010. [DOI] [PubMed] [Google Scholar]
- Jackson AL, Pahl PM, Harrison K, Rosamond J, Sclafani RA. Cell cycle regulation of the yeast Cdc7 protein kinase by association with the Dbf4 protein. Mol Cell Biol. 1993;13:2899–908. doi: 10.1128/mcb.13.5.2899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jares P, Blow JJ. Xenopus cdc7 function is dependent on licensing but not on XORC, XCdc6, or CDK activity and is required for XCdc45 loading. Genes Dev. 2000;14:1528–40. [PMC free article] [PubMed] [Google Scholar]
- Jazwinski SM, Edelman GM. Protein complexes from active replicative fractions associate in vitro with the replication origins of yeast 2-micrometers DNA plasmid. Proc Natl Acad Sci U S A. 1982;79:3428–32. doi: 10.1073/pnas.79.11.3428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jee J, Mizuno T, Kamada K, Tochio H, Chiba Y, Yanagi K, Yasuda G, Hiroaki H, Hanaoka F, Shirakawa M. Structure and mutagenesis studies of the C-terminal region of licensing factor Cdt1 enable the identification of key residues for binding to replicative helicase Mcm proteins. J Biol Chem. 2010;285:15931–40. doi: 10.1074/jbc.M109.075333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jenkinson ER, Chong JP. Minichromosome maintenance helicase activity is controlled by N- and C-terminal motifs and requires the ATPase domain helix-2 insert. Proc Natl Acad Sci U S A. 2006;103:7613–8. doi: 10.1073/pnas.0509297103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang W, Mcdonald D, Hope TJ, Hunter T. Mammalian Cdc7-Dbf4 protein kinase complex is essential for initiation of DNA replication. EMBO J. 1999a;18:5703–13. doi: 10.1093/emboj/18.20.5703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang W, Wells NJ, Hunter T. Multistep regulation of DNA replication by Cdk phosphorylation of HsCdc6. Proc Natl Acad Sci U S A. 1999b;96:6193–8. doi: 10.1073/pnas.96.11.6193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kalfalah FM, Berg E, Christensen MO, Linka RM, Dirks WG, Boege F, Mielke C. Spatio-temporal regulation of the human licensing factor Cdc6 in replication and mitosis. Cell Cycle. 2015;14:1704–15. doi: 10.1080/15384101.2014.1000182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kamimura Y, Masumoto H, Sugino A, Araki H. Sld2, which interacts with Dpb11 in Saccharomyces cerevisiae, is required for chromosomal DNA replication. Mol Cell Biol. 1998;18:6102–9. doi: 10.1128/mcb.18.10.6102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kamimura Y, Tak YS, Sugino A, Araki H. Sld3, which interacts with Cdc45 (Sld4), functions for chromosomal DNA replication in Saccharomyces cerevisiae. EMBO J. 2001;20:2097–107. doi: 10.1093/emboj/20.8.2097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanemaki M, Sanchez-Diaz A, Gambus A, Labib K. Functional proteomic identification of DNA replication proteins by induced proteolysis in vivo. Nature. 2003;423:720–4. doi: 10.1038/nature01692. [DOI] [PubMed] [Google Scholar]
- Kang S, Warner MD, Bell SP. Multiple functions for Mcm2–7 ATPase motifs during replication initiation. Mol Cell. 2014;55:655–65. doi: 10.1016/j.molcel.2014.06.033. [DOI] [PMC free article] [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;109:6042–7. doi: 10.1073/pnas.1203734109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaplan DL. The 3′-tail of a forked-duplex sterically determines whether one or two DNA strands pass through the central channel of a replication-fork helicase. J Mol Biol. 2000;301:285–99. doi: 10.1006/jmbi.2000.3965. [DOI] [PubMed] [Google Scholar]
- Kaplan DL, Davey MJ, O’donnell M. Mcm4,6,7 uses a “pump in ring” mechanism to unwind DNA by steric exclusion and actively translocate along a duplex. J Biol Chem. 2003;278:49171–82. doi: 10.1074/jbc.M308074200. [DOI] [PubMed] [Google Scholar]
- Kaplan DL, O’donnell M. DnaB drives DNA branch migration and dislodges proteins while encircling two DNA strands. Mol Cell. 2002;10:647–57. doi: 10.1016/s1097-2765(02)00642-1. [DOI] [PubMed] [Google Scholar]
- Karnani N, Taylor CM, Malhotra A, Dutta A. Genomic study of replication initiation in human chromosomes reveals the influence of transcription regulation and chromatin structure on origin selection. Mol Biol Cell. 2010;21:393–404. doi: 10.1091/mbc.E09-08-0707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kasiviswanathan R, Shin JH, Melamud E, Kelman Z. Biochemical characterization of the Methanothermobacter thermautotrophicus minichromosome maintenance (MCM) helicase N-terminal domains. J Biol Chem. 2004;279:28358–66. doi: 10.1074/jbc.M403202200. [DOI] [PubMed] [Google Scholar]
- Kawakami H, Ohashi E, Kanamoto S, Tsurimoto T, Katayama T. Specific binding of eukaryotic ORC to DNA replication origins depends on highly conserved basic residues. Sci Rep. 2015;5:14929. doi: 10.1038/srep14929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawasaki Y, Kim HD, Kojima A, Seki T, Sugino A. Reconstitution of Saccharomyces cerevisiae prereplicative complex assembly in vitro. Genes Cells. 2006;11:745–56. doi: 10.1111/j.1365-2443.2006.00975.x. [DOI] [PubMed] [Google Scholar]
- Kelley LA, Mezulis S, Yates CM, Wass MN, Sternberg MJ. The Phyre2 web portal for protein modeling, prediction and analysis. Nat Protoc. 2015;10:845–58. doi: 10.1038/nprot.2015.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kelly TJ, Martin GS, Forsburg SL, Stephen RJ, Russo A, Nurse P. The fission yeast cdc18+ gene product couples S phase to START and mitosis. Cell. 1993;74:371–82. doi: 10.1016/0092-8674(93)90427-r. [DOI] [PubMed] [Google Scholar]
- Kelman Z, Lee JK, Hurwitz J. The single minichromosome maintenance protein of Methanobacterium thermoautotrophicum DeltaH contains DNA helicase activity. Proc Natl Acad Sci U S A. 1999;96:14783–8. doi: 10.1073/pnas.96.26.14783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kerns SL, Torke SJ, Benjamin JM, Mcgarry TJ. Geminin prevents rereplication during xenopus development. J Biol Chem. 2007;282:5514–21. doi: 10.1074/jbc.M609289200. [DOI] [PubMed] [Google Scholar]
- Khayrutdinov BI, Bae WJ, Yun YM, Lee JH, Tsuyama T, Kim JJ, Hwang E, Ryu KS, Cheong HK, Cheong C, Ko JS, Enomoto T, Karplus PA, Guntert P, Tada S, Jeon YH, Cho Y. Structure of the Cdt1 C-terminal domain: conservation of the winged helix fold in replication licensing factors. Protein Sci. 2009;18:2252–64. doi: 10.1002/pro.236. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim SM, Zhang DY, Huberman JA. Multiple redundant sequence elements within the fission yeast ura4 replication origin enhancer. BMC Mol Biol. 2001;2:1. doi: 10.1186/1471-2199-2-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kimura H, Takizawa N, Nozaki N, Sugimoto K. Molecular cloning of cDNA encoding mouse Cdc21 and CDC46 homologs and characterization of the products: physical interaction between P1(MCM3) and CDC46 proteins. Nucleic Acids Res. 1995;23:2097–104. doi: 10.1093/nar/23.12.2097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kitada K, Johnston LH, Sugino T, Sugino A. Temperature-sensitive cdc7 mutations of Saccharomyces cerevisiae are suppressed by the DBF4 gene, which is required for the G1/S cell cycle transition. Genetics. 1992;131:21–9. doi: 10.1093/genetics/131.1.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klemm RD, Austin RJ, Bell SP. Coordinate binding of ATP and origin DNA regulates the ATPase activity of the origin recognition complex. Cell. 1997;88:493–502. doi: 10.1016/s0092-8674(00)81889-9. [DOI] [PubMed] [Google Scholar]
- Knapp D, Bhoite L, Stillman DJ, Nasmyth K. The transcription factor Swi5 regulates expression of the cyclin kinase inhibitor p40SIC1. Mol Cell Biol. 1996;16:5701–7. doi: 10.1128/mcb.16.10.5701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kneissl M, Putter V, Szalay AA, Grummt F. Interaction and assembly of murine pre-replicative complex proteins in yeast and mouse cells. J Mol Biol. 2003;327:111–28. doi: 10.1016/s0022-2836(03)00079-2. [DOI] [PubMed] [Google Scholar]
- Knott SR, Viggiani CJ, Tavare S, Aparicio OM. Genome-wide replication profiles indicate an expansive role for Rpd3L in regulating replication initiation timing or efficiency, and reveal genomic loci of Rpd3 function in Saccharomyces cerevisiae. Genes Dev. 2009;23:1077–90. doi: 10.1101/gad.1784309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kondo T, Kobayashi M, Tanaka J, Yokoyama A, Suzuki S, Kato N, Onozawa M, Chiba K, Hashino S, Imamura M, Minami Y, Minamino N, Asaka M. Rapid degradation of Cdt1 upon UV-induced DNA damage is mediated by SCFSkp2 complex. J Biol Chem. 2004;279:27315–9. doi: 10.1074/jbc.M314023200. [DOI] [PubMed] [Google Scholar]
- Kong D, Coleman TR, Depamphilis ML. Xenopus origin recognition complex (ORC) initiates DNA replication preferentially at sequences targeted by Schizosaccharomyces pombe ORC. EMBO J. 2003;22:3441–50. doi: 10.1093/emboj/cdg319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kong D, Depamphilis ML. Site-specific DNA binding of the Schizosaccharomyces pombe origin recognition complex is determined by the Orc4 subunit. Mol Cell Biol. 2001;21:8095–103. doi: 10.1128/MCB.21.23.8095-8103.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koonin EV. A common set of conserved motifs in a vast variety of putative nucleic acid-dependent ATPases including MCM proteins involved in the initiation of eukaryotic DNA replication. Nucleic Acids Res. 1993;21:2541–7. doi: 10.1093/nar/21.11.2541. [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;287:4121–8. doi: 10.1074/jbc.M111.285395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kreitz S, Ritzi M, Baack M, Knippers R. The human origin recognition complex protein 1 dissociates from chromatin during S phase in HeLa cells. J Biol Chem. 2001;276:6337–42. doi: 10.1074/jbc.M009473200. [DOI] [PubMed] [Google Scholar]
- Krude T, Musahl C, Laskey RA, Knippers R. Human replication proteins hCdc21, hCdc46 and P1Mcm3 bind chromatin uniformly before S-phase and are displaced locally during DNA replication. J Cell Sci. 1996;109(Pt 2):309–18. doi: 10.1242/jcs.109.2.309. [DOI] [PubMed] [Google Scholar]
- Kubota Y, Mimura S, Nishimoto S, Takisawa H, Nojima H. Identification of the yeast MCM3-related protein as a component of Xenopus DNA replication licensing factor. Cell. 1995;81:601–9. doi: 10.1016/0092-8674(95)90081-0. [DOI] [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;17:1141–52. doi: 10.1101/gad.1070003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumagai A, Shevchenko A, Shevchenko A, Dunphy WG. Treslin collaborates with TopBP1 in triggering the initiation of DNA replication. Cell. 2010;140:349–59. doi: 10.1016/j.cell.2009.12.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuo AJ, Song J, Cheung P, Ishibe-Murakami S, Yamazoe S, Chen JK, Patel DJ, Gozani O. The BAH domain of ORC1 links H4K20me2 to DNA replication licensing and Meier-Gorlin syndrome. Nature. 2012;484:115–9. doi: 10.1038/nature10956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Labib K, Diffley JF, Kearsey SE. G1-phase and B-type cyclins exclude the DNA-replication factor Mcm4 from the nucleus. Nat Cell Biol. 1999;1:415–22. doi: 10.1038/15649. [DOI] [PubMed] [Google Scholar]
- Labib K, Tercero JA, Diffley JF. Uninterrupted MCM2–7 function required for DNA replication fork progression. Science. 2000;288:1643–7. doi: 10.1126/science.288.5471.1643. [DOI] [PubMed] [Google Scholar]
- Lam SK, Ma X, Sing TL, Shilton BH, Brandl CJ, Davey MJ. The PS1 hairpin of Mcm3 is essential for viability and for DNA unwinding in vitro. PLoS One. 2013;8:e82177. doi: 10.1371/journal.pone.0082177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Landis G, Kelley R, Spradling AC, Tower J. The k43 gene, required for chorion gene amplification and diploid cell chromosome replication, encodes the Drosophila homolog of yeast origin recognition complex subunit 2. Proc Natl Acad Sci U S A. 1997;94:3888–92. doi: 10.1073/pnas.94.8.3888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leatherwood J, Lopez-Girona A, Russell P. Interaction of Cdc2 and Cdc18 with a fission yeast ORC2-like protein. Nature. 1996;379:360–3. doi: 10.1038/379360a0. [DOI] [PubMed] [Google Scholar]
- Leatherwood J, Vas A. Connecting ORC and heterochromatin: why? Cell Cycle. 2003;2:573–5. [PubMed] [Google Scholar]
- Lee C, Hong B, Choi JM, Kim Y, Watanabe S, Ishimi Y, Enomoto T, Tada S, Kim Y, Cho Y. Structural basis for inhibition of the replication licensing factor Cdt1 by geminin. Nature. 2004;430:913–7. doi: 10.1038/nature02813. [DOI] [PubMed] [Google Scholar]
- Lee DG, Bell SP. Architecture of the yeast origin recognition complex bound to origins of DNA replication. Mol Cell Biol. 1997;17:7159–68. doi: 10.1128/mcb.17.12.7159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee DG, Makhov AM, Klemm RD, Griffith JD, Bell SP. Regulation of origin recognition complex conformation and ATPase activity: differential effects of single-stranded and double-stranded DNA binding. EMBO J. 2000;19:4774–82. doi: 10.1093/emboj/19.17.4774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee JK, Hurwitz J. Isolation and characterization of various complexes of the minichromosome maintenance proteins of Schizosaccharomyces pombe. J Biol Chem. 2000;275:18871–8. doi: 10.1074/jbc.M001118200. [DOI] [PubMed] [Google Scholar]
- Lei M, Kawasaki Y, Young MR, Kihara M, Sugino A, Tye BK. Mcm2 is a target of regulation by Cdc7-Dbf4 during the initiation of DNA synthesis. Genes Dev. 1997;11:3365–74. doi: 10.1101/gad.11.24.3365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leipe DD, Aravind L, Koonin EV. Did DNA replication evolve twice independently? Nucleic Acids Res. 1999;27:3389–401. doi: 10.1093/nar/27.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lenzen CU, Steinmann D, Whiteheart SW, Weis WI. Crystal structure of the hexamerization domain of N-ethylmaleimide-sensitive fusion protein. Cell. 1998;94:525–36. doi: 10.1016/s0092-8674(00)81593-7. [DOI] [PubMed] [Google Scholar]
- Leon RP, Tecklenburg M, Sclafani RA. Functional conservation of beta-hairpin DNA binding domains in the Mcm protein of Methanobacterium thermoautotrophicum and the Mcm5 protein of Saccharomyces cerevisiae. Genetics. 2008;179:1757–68. doi: 10.1534/genetics.108.088690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leonard AC, Mechali M. DNA replication origins. Cold Spring Harb Perspect Biol. 2013;5:a010116. doi: 10.1101/cshperspect.a010116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li CJ, Depamphilis ML. Mammalian Orc1 protein is selectively released from chromatin and ubiquitinated during the S-to-M transition in the cell division cycle. Mol Cell Biol. 2002;22:105–16. doi: 10.1128/MCB.22.1.105-116.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li CJ, Vassilev A, Depamphilis ML. Role for Cdk1 (Cdc2)/cyclin A in preventing the mammalian origin recognition complex’s largest subunit (Orc1) from binding to chromatin during mitosis. Mol Cell Biol. 2004;24:5875–86. doi: 10.1128/MCB.24.13.5875-5886.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li JJ, Herskowitz I. Isolation of ORC6, a component of the yeast origin recognition complex by a one-hybrid system. Science. 1993;262:1870–4. doi: 10.1126/science.8266075. [DOI] [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;524:186–91. doi: 10.1038/nature14685. [DOI] [PubMed] [Google Scholar]
- Li X, Zhao Q, Liao R, Sun P, Wu X. The SCF(Skp2) ubiquitin ligase complex interacts with the human replication licensing factor Cdt1 and regulates Cdt1 degradation. J Biol Chem. 2003;278:30854–8. doi: 10.1074/jbc.C300251200. [DOI] [PubMed] [Google Scholar]
- Liang C, Stillman B. Persistent initiation of DNA replication and chromatin-bound MCM proteins during the cell cycle in cdc6 mutants. Genes Dev. 1997;11:3375–86. doi: 10.1101/gad.11.24.3375. [DOI] [PMC free article] [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;81:667–76. doi: 10.1016/0092-8674(95)90528-6. [DOI] [PubMed] [Google Scholar]
- Lidonnici MR, Rossi R, Paixao S, Mendoza-Maldonado R, Paolinelli R, Arcangeli C, Giacca M, Biamonti G, Montecucco A. Subnuclear distribution of the largest subunit of the human origin recognition complex during the cell cycle. J Cell Sci. 2004;117:5221–31. doi: 10.1242/jcs.01405. [DOI] [PubMed] [Google Scholar]
- Liku ME, Nguyen VQ, Rosales AW, Irie K, Li JJ. CDK phosphorylation of a novel NLS-NES module distributed between two subunits of the Mcm2–7 complex prevents chromosomal rereplication. Mol Biol Cell. 2005;16:5026–39. doi: 10.1091/mbc.E05-05-0412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lipford JR, Bell SP. Nucleosomes positioned by ORC facilitate the initiation of DNA replication. Mol Cell. 2001;7:21–30. doi: 10.1016/s1097-2765(01)00151-4. [DOI] [PubMed] [Google Scholar]
- Lisziewicz J, Godany A, Agoston DV, Kuntzel H. Cloning and characterization of the Saccharomyces cerevisiae CDC6 gene. Nucleic Acids Res. 1988;16:11507–20. doi: 10.1093/nar/16.24.11507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu C, Wu R, Zhou B, Wang J, Wei Z, Tye BK, Liang C, Zhu G. Structural insights into the Cdt1-mediated MCM2–7 chromatin loading. Nucleic Acids Res. 2012a;40:3208–17. doi: 10.1093/nar/gkr1118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu E, Li X, Yan F, Zhao Q, Wu X. Cyclin-dependent kinases phosphorylate human Cdt1 and induce its degradation. J Biol Chem. 2004;279:17283–8. doi: 10.1074/jbc.C300549200. [DOI] [PubMed] [Google Scholar]
- Liu J, Mcconnell K, Dixon M, Calvi BR. Analysis of model replication origins in Drosophila reveals new aspects of the chromatin landscape and its relationship to origin activity and the prereplicative complex. Mol Biol Cell. 2012b;23:200–12. doi: 10.1091/mbc.E11-05-0409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J, Smith CL, Deryckere D, Deangelis K, Martin GS, Berger JM. Structure and function of Cdc6/Cdc18: implications for origin recognition and checkpoint control. Mol Cell. 2000;6:637–48. doi: 10.1016/s1097-2765(00)00062-9. [DOI] [PubMed] [Google Scholar]
- Liu S, Balasov M, Wang H, Wu L, Chesnokov IN, Liu Y. Structural analysis of human Orc6 protein reveals a homology with transcription factor TFIIB. Proc Natl Acad Sci U S A. 2011;108:7373–8. doi: 10.1073/pnas.1013676108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu W, Pucci B, Rossi M, Pisani FM, Ladenstein R. Structural analysis of the Sulfolobus solfataricus MCM protein N-terminal domain. Nucleic Acids Res. 2008;36:3235–43. doi: 10.1093/nar/gkn183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Loo S, Fox CA, Rine J, Kobayashi R, Stillman B, Bell S. The origin recognition complex in silencing, cell cycle progression, and DNA replication. Mol Biol Cell. 1995;6:741–56. doi: 10.1091/mbc.6.6.741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lubelsky Y, Sasaki T, Kuipers MA, Lucas I, Le Beau MM, Carignon S, Debatisse M, Prinz JA, Dennis JH, Gilbert DM. Pre-replication complex proteins assemble at regions of low nucleosome occupancy within the Chinese hamster dihydrofolate reductase initiation zone. Nucleic Acids Res. 2011;39:3141–55. doi: 10.1093/nar/gkq1276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lutzmann M, Maiorano D, Mechali M. A Cdt1-geminin complex licenses chromatin for DNA replication and prevents rereplication during S phase in Xenopus. EMBO J. 2006;25:5764–74. doi: 10.1038/sj.emboj.7601436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lyubimov AY, Costa A, Bleichert F, Botchan MR, Berger JM. ATP-dependent conformational dynamics underlie the functional asymmetry of the replicative helicase from a minimalist eukaryote. Proc Natl Acad Sci U S A. 2012;109:11999–2004. doi: 10.1073/pnas.1209406109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Macalpine DM, Bell SP. A genomic view of eukaryotic DNA replication. Chromosome Res. 2005;13:309–26. doi: 10.1007/s10577-005-1508-1. [DOI] [PubMed] [Google Scholar]
- Macalpine HK, Gordan R, Powell SK, Hartemink AJ, Macalpine DM. Drosophila ORC localizes to open chromatin and marks sites of cohesin complex loading. Genome Res. 2010;20:201–11. doi: 10.1101/gr.097873.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Madine MA, Khoo CY, Mills AD, Laskey RA. MCM3 complex required for cell cycle regulation of DNA replication in vertebrate cells. Nature. 1995;375:421–4. doi: 10.1038/375421a0. [DOI] [PubMed] [Google Scholar]
- Maine GT, Sinha P, Tye BK. Mutants of S. cerevisiae defective in the maintenance of minichromosomes. Genetics. 1984;106:365–85. doi: 10.1093/genetics/106.3.365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maiorano D, Moreau J, Mechali M. XCDT1 is required for the assembly of pre-replicative complexes in Xenopus laevis. Nature. 2000;404:622–5. doi: 10.1038/35007104. [DOI] [PubMed] [Google Scholar]
- Maiorano D, Rul W, Mechali M. Cell cycle regulation of the licensing activity of Cdt1 in Xenopus laevis. Exp Cell Res. 2004;295:138–49. doi: 10.1016/j.yexcr.2003.11.018. [DOI] [PubMed] [Google Scholar]
- Makarova KS, Koonin EV, Kelman Z. The CMG (CDC45/RecJ, MCM, GINS) complex is a conserved component of the DNA replication system in all archaea and eukaryotes. Biol Direct. 2012;7:7. doi: 10.1186/1745-6150-7-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Makise M, Takenaka H, Kuwae W, Takahashi N, Tsuchiya T, Mizushima T. Kinetics of ATP binding to the origin recognition complex of Saccharomyces cerevisiae. J Biol Chem. 2003;278:46440–5. doi: 10.1074/jbc.M307392200. [DOI] [PubMed] [Google Scholar]
- Marahrens Y, Stillman B. A yeast chromosomal origin of DNA replication defined by multiple functional elements. Science. 1992;255:817–23. doi: 10.1126/science.1536007. [DOI] [PubMed] [Google Scholar]
- Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N, Oda M. Eukaryotic chromosome DNA replication: where, when, and how? Annu Rev Biochem. 2010;79:89–130. doi: 10.1146/annurev.biochem.052308.103205. [DOI] [PubMed] [Google Scholar]
- Masai H, Taniyama C, Ogino K, Matsui E, Kakusho N, Matsumoto S, Kim JM, Ishii A, Tanaka T, Kobayashi T, Tamai K, Ohtani K, Arai K. Phosphorylation of MCM4 by Cdc7 kinase facilitates its interaction with Cdc45 on the chromatin. J Biol Chem. 2006;281:39249–61. doi: 10.1074/jbc.M608935200. [DOI] [PubMed] [Google Scholar]
- Masuda T, Mimura S, Takisawa H. CDK- and Cdc45-dependent priming of the MCM complex on chromatin during S-phase in Xenopus egg extracts: possible activation of MCM helicase by association with Cdc45. Genes Cells. 2003;8:145–61. doi: 10.1046/j.1365-2443.2003.00621.x. [DOI] [PubMed] [Google Scholar]
- Masumoto H, Muramatsu S, Kamimura Y, Araki H. S-Cdk-dependent phosphorylation of Sld2 essential for chromosomal DNA replication in budding yeast. Nature. 2002;415:651–5. doi: 10.1038/nature713. [DOI] [PubMed] [Google Scholar]
- Matsuno K, Kumano M, Kubota Y, Hashimoto Y, Takisawa H. The N-terminal noncatalytic region of Xenopus RecQ4 is required for chromatin binding of DNA polymerase alpha in the initiation of DNA replication. Mol Cell Biol. 2006;26:4843–52. doi: 10.1128/MCB.02267-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mcgarry TJ, Kirschner MW. Geminin, an inhibitor of DNA replication, is degraded during mitosis. Cell. 1998;93:1043–53. doi: 10.1016/s0092-8674(00)81209-x. [DOI] [PubMed] [Google Scholar]
- Mcgeoch AT, Trakselis MA, Laskey RA, Bell SD. Organization of the archaeal MCM complex on DNA and implications for the helicase mechanism. Nat Struct Mol Biol. 2005;12:756–62. doi: 10.1038/nsmb974. [DOI] [PubMed] [Google Scholar]
- Mechali M, Kearsey S. Lack of specific sequence requirement for DNA replication in Xenopus eggs compared with high sequence specificity in yeast. Cell. 1984;38:55–64. doi: 10.1016/0092-8674(84)90526-9. [DOI] [PubMed] [Google Scholar]
- Mendenhall MD. An inhibitor of p34CDC28 protein kinase activity from Saccharomyces cerevisiae. Science. 1993;259:216–9. doi: 10.1126/science.8421781. [DOI] [PubMed] [Google Scholar]
- Mendez J, Zou-Yang XH, Kim SY, Hidaka M, Tansey WP, Stillman B. Human origin recognition complex large subunit is degraded by ubiquitin-mediated proteolysis after initiation of DNA replication. Mol Cell. 2002;9:481–91. doi: 10.1016/s1097-2765(02)00467-7. [DOI] [PubMed] [Google Scholar]
- Micklem G, Rowley A, Harwood J, Nasmyth K, Diffley JF. Yeast origin recognition complex is involved in DNA replication and transcriptional silencing. Nature. 1993;366:87–9. doi: 10.1038/366087a0. [DOI] [PubMed] [Google Scholar]
- Mihaylov IS, Kondo T, Jones L, Ryzhikov S, Tanaka J, Zheng J, Higa LA, Minamino N, Cooley L, Zhang H. Control of DNA replication and chromosome ploidy by geminin and cyclin A. Mol Cell Biol. 2002;22:1868–80. doi: 10.1128/MCB.22.6.1868-1880.2002. [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;431:1118–23. doi: 10.1038/nature03024. [DOI] [PubMed] [Google Scholar]
- Miotto B, Struhl K. HBO1 histone acetylase activity is essential for DNA replication licensing and inhibited by Geminin. Mol Cell. 2010;37:57–66. doi: 10.1016/j.molcel.2009.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mizushima T, Takahashi N, Stillman B. Cdc6p modulates the structure and DNA binding activity of the origin recognition complex in vitro. Genes Dev. 2000;14:1631–41. [PMC free article] [PubMed] [Google Scholar]
- Moir D, Stewart SE, Osmond BC, Botstein D. Cold-sensitive cell-division-cycle mutants of yeast: isolation, properties, and pseudoreversion studies. Genetics. 1982;100:547–63. doi: 10.1093/genetics/100.4.547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moll T, Tebb G, Surana U, Robitsch H, Nasmyth K. The role of phosphorylation and the CDC28 protein kinase in cell cycle-regulated nuclear import of the S. cerevisiae transcription factor SWI5. Cell. 1991;66:743–58. doi: 10.1016/0092-8674(91)90118-i. [DOI] [PubMed] [Google Scholar]
- Montagnoli A, Valsasina B, Brotherton D, Troiani S, Rainoldi S, Tenca P, Molinari A, Santocanale C. Identification of Mcm2 phosphorylation sites by S-phase-regulating kinases. J Biol Chem. 2006;281:10281–90. doi: 10.1074/jbc.M512921200. [DOI] [PubMed] [Google Scholar]
- Moon KY, Kong D, Lee JK, Raychaudhuri S, Hurwitz J. Identification and reconstitution of the origin recognition complex from Schizosaccharomyces pombe. Proc Natl Acad Sci U S A. 1999;96:12367–72. doi: 10.1073/pnas.96.22.12367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moreau MJ, Mcgeoch AT, Lowe AR, Itzhaki LS, Bell SD. ATPase site architecture and helicase mechanism of an archaeal MCM. Mol Cell. 2007;28:304–14. doi: 10.1016/j.molcel.2007.08.013. [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;103:10236–41. doi: 10.1073/pnas.0602400103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muller P, Park S, Shor E, Huebert DJ, Warren CL, Ansari AZ, Weinreich M, Eaton ML, Macalpine DM, Fox CA. The conserved bromo-adjacent homology domain of yeast Orc1 functions in the selection of DNA replication origins within chromatin. Genes Dev. 2010;24:1418–33. doi: 10.1101/gad.1906410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muramatsu S, Hirai K, Tak YS, Kamimura Y, Araki H. CDK-dependent complex formation between replication proteins Dpb11, Sld2, Pol (epsilon}, and GINS in budding yeast. Genes Dev. 2010;24:602–12. doi: 10.1101/gad.1883410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Musahl C, Schulte D, Burkhart R, Knippers R. A human homologue of the yeast replication protein Cdc21. Interactions with other Mcm proteins. Eur J Biochem. 1995;230:1096–101. doi: 10.1111/j.1432-1033.1995.tb20660.x. [DOI] [PubMed] [Google Scholar]
- Muzi-Falconi M, Kelly TJ. Orp1, a member of the Cdc18/Cdc6 family of S-phase regulators, is homologous to a component of the origin recognition complex. Proc Natl Acad Sci U S A. 1995;92:12475–9. doi: 10.1073/pnas.92.26.12475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Narasimhachar Y, Coue M. Geminin stabilizes Cdt1 during meiosis in Xenopus oocytes. J Biol Chem. 2009;284:27235–42. doi: 10.1074/jbc.M109.008854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Natale DA, Li CJ, Sun WH, Depamphilis ML. Selective instability of Orc1 protein accounts for the absence of functional origin recognition complexes during the M-G(1) transition in mammals. EMBO J. 2000;19:2728–38. doi: 10.1093/emboj/19.11.2728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neuwald AF, Aravind L, Spouge JL, Koonin EV. AAA+: A class of chaperone-like ATPases associated with the assembly, operation, and disassembly of protein complexes. Genome Res. 1999;9:27–43. [PubMed] [Google Scholar]
- Nguyen VQ, Co C, Li JJ. Cyclin-dependent kinases prevent DNA re-replication through multiple mechanisms. Nature. 2001;411:1068–73. doi: 10.1038/35082600. [DOI] [PubMed] [Google Scholar]
- Nieduszynski CA, Knox Y, Donaldson AD. Genome-wide identification of replication origins in yeast by comparative genomics. Genes Dev. 2006;20:1874–9. doi: 10.1101/gad.385306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nishitani H, Lygerou Z, Nishimoto T. Proteolysis of DNA replication licensing factor Cdt1 in S-phase is performed independently of geminin through its N-terminal region. J Biol Chem. 2004;279:30807–16. doi: 10.1074/jbc.M312644200. [DOI] [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;404:625–8. doi: 10.1038/35007110. [DOI] [PubMed] [Google Scholar]
- Nishitani H, Sugimoto N, Roukos V, Nakanishi Y, Saijo M, Obuse C, Tsurimoto T, Nakayama KI, Nakayama K, Fujita M, Lygerou Z, Nishimoto T. Two E3 ubiquitin ligases, SCF-Skp2 and DDB1-Cul4, target human Cdt1 for proteolysis. EMBO J. 2006;25:1126–36. doi: 10.1038/sj.emboj.7601002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nishitani H, Taraviras S, Lygerou Z, Nishimoto T. The human licensing factor for DNA replication Cdt1 accumulates in G1 and is destabilized after initiation of S-phase. J Biol Chem. 2001;276:44905–11. doi: 10.1074/jbc.M105406200. [DOI] [PubMed] [Google Scholar]
- Noguchi K, Vassilev A, Ghosh S, Yates JL, Depamphilis ML. The BAH domain facilitates the ability of human Orc1 protein to activate replication origins in vivo. EMBO J. 2006;25:5372–82. doi: 10.1038/sj.emboj.7601396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nougarede R, Della Seta F, Zarzov P, Schwob E. Hierarchy of S-phase-promoting factors: yeast Dbf4-Cdc7 kinase requires prior S-phase cyclin-dependent kinase activation. Mol Cell Biol. 2000;20:3795–806. doi: 10.1128/mcb.20.11.3795-3806.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohlenschlager O, Kuhnert A, Schneider A, Haumann S, Bellstedt P, Keller H, Saluz HP, Hortschansky P, Hanel F, Grosse F, Gorlach M, Pospiech H. The N-terminus of the human RecQL4 helicase is a homeodomain-like DNA interaction motif. Nucleic Acids Res. 2012;40:8309–24. doi: 10.1093/nar/gks591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohta S, Tatsumi Y, Fujita M, Tsurimoto T, Obuse C. The ORC1 cycle in human cells: II. Dynamic changes in the human ORC complex during the cell cycle. J Biol Chem. 2003;278:41535–40. doi: 10.1074/jbc.M307535200. [DOI] [PubMed] [Google Scholar]
- Okuno Y, Satoh H, Sekiguchi M, Masukata H. Clustered adenine/thymine stretches are essential for function of a fission yeast replication origin. Mol Cell Biol. 1999;19:6699–709. doi: 10.1128/mcb.19.10.6699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oshiro G, Owens JC, Shellman Y, Sclafani RA, Li JJ. Cell cycle control of Cdc7p kinase activity through regulation of Dbf4p stability. Mol Cell Biol. 1999;19:4888–96. doi: 10.1128/mcb.19.7.4888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pacek M, Tutter AV, Kubota Y, Takisawa H, Walter JC. Localization of MCM2–7, Cdc45, and GINS to the site of DNA unwinding during eukaryotic DNA replication. Mol Cell. 2006;21:581–7. doi: 10.1016/j.molcel.2006.01.030. [DOI] [PubMed] [Google Scholar]
- Pacek M, Walter JC. A requirement for MCM7 and Cdc45 in chromosome unwinding during eukaryotic DNA replication. EMBO J. 2004;23:3667–76. doi: 10.1038/sj.emboj.7600369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pak DT, Pflumm M, Chesnokov I, Huang DW, Kellum R, Marr J, Romanowski P, Botchan MR. Association of the origin recognition complex with heterochromatin and HP1 in higher eukaryotes. Cell. 1997;91:311–23. doi: 10.1016/s0092-8674(00)80415-8. [DOI] [PubMed] [Google Scholar]
- Palacios Debeer MA, Muller U, Fox CA. Differential DNA affinity specifies roles for the origin recognition complex in budding yeast heterochromatin. Genes Dev. 2003;17:1817–22. doi: 10.1101/gad.1096703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palmer RE, Hogan E, Koshland D. Mitotic transmission of artificial chromosomes in cdc mutants of the yeast, Saccharomyces cerevisiae. Genetics. 1990;125:763–74. doi: 10.1093/genetics/125.4.763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perez-Arnaiz P, Kaplan DL. An Mcm10 Mutant Defective in ssDNA Binding Shows Defects in DNA Replication Initiation. J Mol Biol. 2016;428:4608–4625. doi: 10.1016/j.jmb.2016.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perkins G, Diffley JF. Nucleotide-dependent prereplicative complex assembly by Cdc6p, a homolog of eukaryotic and prokaryotic clamp-loaders. Mol Cell. 1998;2:23–32. doi: 10.1016/s1097-2765(00)80110-0. [DOI] [PubMed] [Google Scholar]
- Perkins G, Drury LS, Diffley JF. Separate SCF(CDC4) recognition elements target Cdc6 for proteolysis in S phase and mitosis. EMBO J. 2001;20:4836–45. doi: 10.1093/emboj/20.17.4836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Petersen BO, Lukas J, Sorensen CS, Bartek J, Helin K. Phosphorylation of mammalian CDC6 by cyclin A/CDK2 regulates its subcellular localization. EMBO J. 1999;18:396–410. doi: 10.1093/emboj/18.2.396. [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:E249–58. doi: 10.1073/pnas.1422003112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piatti S, Lengauer C, Nasmyth K. Cdc6 is an unstable protein whose de novo synthesis in G1 is important for the onset of S phase and for preventing a ‘reductional’ anaphase in the budding yeast Saccharomyces cerevisiae. EMBO J. 1995;14:3788–99. doi: 10.1002/j.1460-2075.1995.tb00048.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poplawski A, Grabowski B, Long SE, Kelman Z. The zinc finger domain of the archaeal minichromosome maintenance protein is required for helicase activity. J Biol Chem. 2001;276:49371–7. doi: 10.1074/jbc.M108519200. [DOI] [PubMed] [Google Scholar]
- Powell SK, Macalpine HK, Prinz JA, Li YL, Belsky JA, Macalpine DM. Dynamic loading and redistribution of the Mcm2–7 helicase complex through the cell cycle. Embo Journal. 2015;34:531–543. doi: 10.15252/embj.201488307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasanth SG, Prasanth KV, Siddiqui K, Spector DL, Stillman B. Human Orc2 localizes to centrosomes, centromeres and heterochromatin during chromosome inheritance. EMBO J. 2004;23:2651–63. doi: 10.1038/sj.emboj.7600255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasanth SG, Shen Z, Prasanth KV, Stillman B. Human origin recognition complex is essential for HP1 binding to chromatin and heterochromatin organization. Proc Natl Acad Sci U S A. 2010;107:15093–8. doi: 10.1073/pnas.1009945107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prioleau MN. CpG islands: starting blocks for replication and transcription. PLoS Genet. 2009;5:e1000454. doi: 10.1371/journal.pgen.1000454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prokhorova TA, Blow JJ. Sequential MCM/P1 subcomplex assembly is required to form a heterohexamer with replication licensing activity. J Biol Chem. 2000;275:2491–8. doi: 10.1074/jbc.275.4.2491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pucci B, De Felice M, Rocco M, Esposito F, De Falco M, Esposito L, Rossi M, Pisani FM. Modular organization of the Sulfolobus solfataricus mini-chromosome maintenance protein. J Biol Chem. 2007;282:12574–82. doi: 10.1074/jbc.M610953200. [DOI] [PubMed] [Google Scholar]
- Pucci B, De Felice M, Rossi M, Onesti S, Pisani FM. Amino acids of the Sulfolobus solfataricus mini-chromosome maintenance-like DNA helicase involved in DNA binding/remodeling. J Biol Chem. 2004;279:49222–8. doi: 10.1074/jbc.M408967200. [DOI] [PubMed] [Google Scholar]
- Putnam CD, Clancy SB, Tsuruta H, Gonzalez S, Wetmur JG, Tainer JA. Structure and mechanism of the RuvB Holliday junction branch migration motor. J Mol Biol. 2001;311:297–310. doi: 10.1006/jmbi.2001.4852. [DOI] [PubMed] [Google Scholar]
- Quinn LM, Herr A, Mcgarry TJ, Richardson H. The Drosophila Geminin homolog: roles for Geminin in limiting DNA replication, in anaphase and in neurogenesis. Genes Dev. 2001;15:2741–54. doi: 10.1101/gad.916201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quintana DG, Hou Z, Thome KC, Hendricks M, Saha P, Dutta A. Identification of HsORC4, a member of the human origin of replication recognition complex. J Biol Chem. 1997;272:28247–51. doi: 10.1074/jbc.272.45.28247. [DOI] [PubMed] [Google Scholar]
- Ramer MD, Suman ES, Richter H, Stanger K, Spranger M, Bieberstein N, Duncker BP. Dbf4 and Cdc7 proteins promote DNA replication through interactions with distinct Mcm2–7 protein subunits. J Biol Chem. 2013;288:14926–35. doi: 10.1074/jbc.M112.392910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramey CJ, Sclafani RA. Functional conservation of the pre-sensor one beta-finger hairpin (PS1-hp) structures in mini-chromosome maintenance proteins of Saccharomyces cerevisiae and archaea. G3 (Bethesda) 2014;4:1319–26. doi: 10.1534/g3.114.011668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Randell JC, Bowers JL, Rodriguez HK, Bell SP. Sequential ATP hydrolysis by Cdc6 and ORC directs loading of the Mcm2–7 helicase. Mol Cell. 2006;21:29–39. doi: 10.1016/j.molcel.2005.11.023. [DOI] [PubMed] [Google Scholar]
- Randell JC, Fan A, Chan C, Francis LI, Heller RC, Galani K, Bell SP. Mec1 is one of multiple kinases that prime the Mcm2–7 helicase for phosphorylation by Cdc7. Mol Cell. 2010;40:353–63. doi: 10.1016/j.molcel.2010.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ranjan A, Gossen M. A structural role for ATP in the formation and stability of the human origin recognition complex. Proc Natl Acad Sci U S A. 2006;103:4864–9. doi: 10.1073/pnas.0510305103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rao H, Marahrens Y, Stillman B. Functional conservation of multiple elements in yeast chromosomal replicators. Mol Cell Biol. 1994;14:7643–51. doi: 10.1128/mcb.14.11.7643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rao H, Stillman B. The origin recognition complex interacts with a bipartite DNA binding site within yeast replicators. Proc Natl Acad Sci U S A. 1995;92:2224–8. doi: 10.1073/pnas.92.6.2224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Remus D, Beall EL, Botchan MR. DNA topology, not DNA sequence, is a critical determinant for Drosophila ORC-DNA binding. EMBO J. 2004;23:897–907. doi: 10.1038/sj.emboj.7600077. [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;139:719–30. doi: 10.1016/j.cell.2009.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rice JC, Nishioka K, Sarma K, Steward R, Reinberg D, Allis CD. Mitotic-specific methylation of histone H4 Lys 20 follows increased PR-Set7 expression and its localization to mitotic chromosomes. Genes Dev. 2002;16:2225–30. doi: 10.1101/gad.1014902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ricke RM, Bielinsky AK. Mcm10 regulates the stability and chromatin association of DNA polymerase-alpha. Mol Cell. 2004;16:173–85. doi: 10.1016/j.molcel.2004.09.017. [DOI] [PubMed] [Google Scholar]
- Ritzi M, Baack M, Musahl C, Romanowski P, Laskey RA, Knippers R. Human minichromosome maintenance proteins and human origin recognition complex 2 protein oil chromatin. Journal of Biological Chemistry. 1998;273:24543–24549. doi: 10.1074/jbc.273.38.24543. [DOI] [PubMed] [Google Scholar]
- Robinson NP, Bell SD. Extrachromosomal element capture and the evolution of multiple replication origins in archaeal chromosomes. Proc Natl Acad Sci U S A. 2007;104:5806–11. doi: 10.1073/pnas.0700206104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romanowski P, Madine MA, Laskey RA. XMCM7, a novel member of the Xenopus MCM family, interacts with XMCM3 and colocalizes with it throughout replication. Proc Natl Acad Sci U S A. 1996a;93:10189–94. doi: 10.1073/pnas.93.19.10189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romanowski P, Madine MA, Rowles A, Blow JJ, Laskey RA. The Xenopus origin recognition complex is essential for DNA replication and MCM binding to chromatin. Curr Biol. 1996b;6:1416–25. doi: 10.1016/s0960-9822(96)00746-4. [DOI] [PubMed] [Google Scholar]
- Rothenberg E, Trakselis MA, Bell SD, Ha T. MCM forked substrate specificity involves dynamic interaction with the 5′-tail. J Biol Chem. 2007;282:34229–34. doi: 10.1074/jbc.M706300200. [DOI] [PubMed] [Google Scholar]
- Rowles A, Chong JP, Brown L, Howell M, Evan GI, Blow JJ. Interaction between the origin recognition complex and the replication licensing system in Xenopus. Cell. 1996;87:287–96. doi: 10.1016/s0092-8674(00)81346-x. [DOI] [PubMed] [Google Scholar]
- Rowles A, Tada S, Blow JJ. Changes in association of the Xenopus origin recognition complex with chromatin on licensing of replication origins. J Cell Sci. 1999;112(Pt 12):2011–8. doi: 10.1242/jcs.112.12.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Royzman I, Austin RJ, Bosco G, Bell SP, Orr-Weaver TL. ORC localization in Drosophila follicle cells and the effects of mutations in dE2F and dDP. Genes Dev. 1999;13:827–40. doi: 10.1101/gad.13.7.827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saha P, Chen J, Thome KC, Lawlis SJ, Hou ZH, Hendricks M, Parvin JD, Dutta A. Human CDC6/Cdc18 associates with Orc1 and cyclin-cdk and is selectively eliminated from the nucleus at the onset of S phase. Mol Cell Biol. 1998;18:2758–67. doi: 10.1128/mcb.18.5.2758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sakakibara N, Kasiviswanathan R, Melamud E, Han M, Schwarz FP, Kelman Z. Coupling of DNA binding and helicase activity is mediated by a conserved loop in the MCM protein. Nucleic Acids Res. 2008;36:1309–20. doi: 10.1093/nar/gkm1160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Samel SA, Fernandez-Cid A, Sun J, Riera A, Tognetti S, Herrera MC, Li H, Speck C. A unique DNA entry gate serves for regulated loading of the eukaryotic replicative helicase MCM2–7 onto DNA. Genes Dev. 2014;28:1653–66. doi: 10.1101/gad.242404.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Samson RY, Abeyrathne PD, Bell SD. Mechanism of Archaeal MCM Helicase Recruitment to DNA Replication Origins. Mol Cell. 2016;61:287–96. doi: 10.1016/j.molcel.2015.12.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Samson RY, Xu Y, Gadelha C, Stone TA, Faqiri JN, Li D, Qin N, Pu F, Liang YX, She Q, Bell SD. Specificity and function of archaeal DNA replication initiator proteins. Cell Rep. 2013;3:485–96. doi: 10.1016/j.celrep.2013.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sanchez-Pulido L, Diffley JF, Ponting CP. Homology explains the functional similarities of Treslin/Ticrr and Sld3. Curr Biol. 2010;20:R509–10. doi: 10.1016/j.cub.2010.05.021. [DOI] [PubMed] [Google Scholar]
- Sanchez-Pulido L, Ponting CP. Cdc45: the missing RecJ ortholog in eukaryotes? Bioinformatics. 2011;27:1885–8. doi: 10.1093/bioinformatics/btr332. [DOI] [PubMed] [Google Scholar]
- Sangrithi MN, Bernal JA, Madine M, Philpott A, Lee J, Dunphy WG, Venkitaraman AR. Initiation of DNA replication requires the RECQL4 protein mutated in Rothmund-Thomson syndrome. Cell. 2005;121:887–98. doi: 10.1016/j.cell.2005.05.015. [DOI] [PubMed] [Google Scholar]
- Sansam CL, Cruz NM, Danielian PS, Amsterdam A, Lau ML, Hopkins N, Lees JA. A vertebrate gene, ticrr, is an essential checkpoint and replication regulator. Genes Dev. 2010;24:183–94. doi: 10.1101/gad.1860310. [DOI] [PMC free article] [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;15:6671–9. [PMC free article] [PubMed] [Google Scholar]
- Saxena S, Yuan P, Dhar SK, Senga T, Takeda D, Robinson H, Kornbluth S, Swaminathan K, Dutta A. A dimerized coiled-coil domain and an adjoining part of geminin interact with two sites on Cdt1 for replication inhibition. Mol Cell. 2004;15:245–58. doi: 10.1016/j.molcel.2004.06.045. [DOI] [PubMed] [Google Scholar]
- Schepers A, Diffley JF. Mutational analysis of conserved sequence motifs in the budding yeast Cdc6 protein. J Mol Biol. 2001;308:597–608. doi: 10.1006/jmbi.2001.4637. [DOI] [PubMed] [Google Scholar]
- Schneider BL, Yang QH, Futcher AB. Linkage of replication to start by the Cdk inhibitor Sic1. Science. 1996;272:560–2. doi: 10.1126/science.272.5261.560. [DOI] [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;8:1093–104. doi: 10.1016/s1097-2765(01)00389-6. [DOI] [PubMed] [Google Scholar]
- Schwob E, Bohm T, Mendenhall MD, Nasmyth K. The B-type cyclin kinase inhibitor p40SIC1 controls the G1 to S transition in S. cerevisiae. Cell. 1994;79:233–44. doi: 10.1016/0092-8674(94)90193-7. [DOI] [PubMed] [Google Scholar]
- Segurado M, De Luis A, Antequera F. Genome-wide distribution of DNA replication origins at A+T-rich islands in Schizosaccharomyces pombe. EMBO Rep. 2003;4:1048–53. doi: 10.1038/sj.embor.7400008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seki T, Diffley JF. Stepwise assembly of initiation proteins at budding yeast replication origins in vitro. Proc Natl Acad Sci U S A. 2000;97:14115–20. doi: 10.1073/pnas.97.26.14115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Semple JW, Da-Silva LF, Jervis EJ, Ah-Kee J, Al-Attar H, Kummer L, Heikkila JJ, Pasero P, Duncker BP. An essential role for Orc6 in DNA replication through maintenance of pre-replicative complexes. EMBO J. 2006;25:5150–8. doi: 10.1038/sj.emboj.7601391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Senga T, Sivaprasad U, Zhu W, Park JH, Arias EE, Walter JC, Dutta A. PCNA is a cofactor for Cdt1 degradation by CUL4/DDB1-mediated N-terminal ubiquitination. J Biol Chem. 2006;281:6246–52. doi: 10.1074/jbc.M512705200. [DOI] [PubMed] [Google Scholar]
- Sequeira-Mendes J, Diaz-Uriarte R, Apedaile A, Huntley D, Brockdorff N, Gomez M. Transcription initiation activity sets replication origin efficiency in mammalian cells. PLoS Genet. 2009;5:e1000446. doi: 10.1371/journal.pgen.1000446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shechter D, Ying CY, Gautier J. DNA unwinding is an Mcm complex-dependent and ATP hydrolysis-dependent process. J Biol Chem. 2004;279:45586–93. doi: 10.1074/jbc.M407772200. [DOI] [PubMed] [Google Scholar]
- Shechter DF, Ying CY, Gautier J. The intrinsic DNA helicase activity of Methanobacterium thermoautotrophicum delta H minichromosome maintenance protein. J Biol Chem. 2000;275:15049–59. doi: 10.1074/jbc.M000398200. [DOI] [PubMed] [Google Scholar]
- Shen Z, Chakraborty A, Jain A, Giri S, Ha T, Prasanth KV, Prasanth SG. Dynamic association of ORCA with prereplicative complex components regulates DNA replication initiation. Mol Cell Biol. 2012;32:3107–20. doi: 10.1128/MCB.00362-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shen Z, Sathyan KM, Geng Y, Zheng R, Chakraborty A, Freeman B, Wang F, Prasanth KV, Prasanth SG. A WD-repeat protein stabilizes ORC binding to chromatin. Mol Cell. 2010;40:99–111. doi: 10.1016/j.molcel.2010.09.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheu YJ, Stillman B. Cdc7-Dbf4 phosphorylates MCM proteins via a docking site-mediated mechanism to promote S phase progression. Mol Cell. 2006;24:101–13. doi: 10.1016/j.molcel.2006.07.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheu YJ, Stillman B. The Dbf4-Cdc7 kinase promotes S phase by alleviating an inhibitory activity in Mcm4. Nature. 2010;463:113–7. doi: 10.1038/nature08647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shor E, Warren CL, Tietjen J, Hou Z, Muller U, Alborelli I, Gohard FH, Yemm AI, Borisov L, Broach JR, Weinreich M, Nieduszynski CA, Ansari AZ, Fox CA. The origin recognition complex interacts with a subset of metabolic genes tightly linked to origins of replication. PLoS Genet. 2009;5:e1000755. doi: 10.1371/journal.pgen.1000755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shore D, Stillman DJ, Brand AH, Nasmyth KA. Identification of silencer binding proteins from yeast: possible roles in SIR control and DNA replication. EMBO J. 1987;6:461–7. doi: 10.1002/j.1460-2075.1987.tb04776.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Siddiqui K, On KF, Diffley JF. Regulating DNA replication in eukarya. Cold Spring Harb Perspect Biol. 2013:5. doi: 10.1101/cshperspect.a012930. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Siddiqui K, Stillman B. ATP-dependent assembly of the human origin recognition complex. J Biol Chem. 2007;282:32370–83. doi: 10.1074/jbc.M705905200. [DOI] [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]
- Simon AC, Zhou JC, Perera RL, Van Deursen F, Evrin C, Ivanova ME, Kilkenny ML, Renault L, Kjaer S, Matak-Vinkovic D, Labib K, Costa A, Pellegrini L. A Ctf4 trimer couples the CMG helicase to DNA polymerase alpha in the eukaryotic replisome. Nature. 2014;510:293–7. doi: 10.1038/nature13234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simpson RT. Nucleosome positioning can affect the function of a cis-acting DNA element in vivo. Nature. 1990;343:387–9. doi: 10.1038/343387a0. [DOI] [PubMed] [Google Scholar]
- Singleton MR, Morales R, Grainge I, Cook N, Isupov MN, Wigley DB. Conformational changes induced by nucleotide binding in Cdc6/ORC from Aeropyrum pernix. J Mol Biol. 2004;343:547–57. doi: 10.1016/j.jmb.2004.08.044. [DOI] [PubMed] [Google Scholar]
- Speck C, Chen Z, Li H, Stillman B. ATPase-dependent cooperative binding of ORC and Cdc6 to origin DNA. Nat Struct Mol Biol. 2005;12:965–71. doi: 10.1038/nsmb1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Speck C, Stillman B. Cdc6 ATPase activity regulates ORC x Cdc6 stability and the selection of specific DNA sequences as origins of DNA replication. J Biol Chem. 2007;282:11705–14. doi: 10.1074/jbc.M700399200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stinchcomb DT, Struhl K, Davis RW. Isolation and characterisation of a yeast chromosomal replicator. Nature. 1979;282:39–43. doi: 10.1038/282039a0. [DOI] [PubMed] [Google Scholar]
- Su TT, Feger G, O’farrell PH. Drosophila MCM protein complexes. Mol Biol Cell. 1996;7:319–29. doi: 10.1091/mbc.7.2.319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sugimoto N, Kitabayashi I, Osano S, Tatsumi Y, Yugawa T, Narisawa-Saito M, Matsukage A, Kiyono T, Fujita M. Identification of novel human Cdt1-binding proteins by a proteomics approach: proteolytic regulation by APC/CCdh1. Mol Biol Cell. 2008;19:1007–21. doi: 10.1091/mbc.E07-09-0859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sugimoto N, Tatsumi Y, Tsurumi T, Matsukage A, Kiyono T, Nishitani H, Fujita M. Cdt1 phosphorylation by cyclin A-dependent kinases negatively regulates its function without affecting geminin binding. J Biol Chem. 2004;279:19691–7. doi: 10.1074/jbc.M313175200. [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;20:944–51. 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;28:2291–303. doi: 10.1101/gad.242313.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun J, Kawakami H, Zech J, Speck C, Stillman B, Li H. Cdc6-induced conformational changes in ORC bound to origin DNA revealed by cryo-electron microscopy. Structure. 2012;20:534–44. doi: 10.1016/j.str.2012.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun WH, Coleman TR, Depamphilis ML. Cell cycle-dependent regulation of the association between origin recognition proteins and somatic cell chromatin. EMBO J. 2002;21:1437–46. doi: 10.1093/emboj/21.6.1437. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sweder KS, Rhode PR, Campbell JL. Purification and characterization of proteins that bind to yeast ARSs. J Biol Chem. 1988;263:17270–7. [PubMed] [Google Scholar]
- Szambowska A, Tessmer I, Kursula P, Usskilat C, Prus P, Pospiech H, Grosse F. DNA binding properties of human Cdc45 suggest a function as molecular wedge for DNA unwinding. Nucleic Acids Res. 2014;42:2308–19. doi: 10.1093/nar/gkt1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tada S, Chong JP, Mahbubani HM, Blow JJ. The RLF-B component of the replication licensing system is distinct from Cdc6 and functions after Cdc6 binds to chromatin. Curr Biol. 1999;9:211–4. doi: 10.1016/s0960-9822(99)80092-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tada S, Li A, Maiorano D, Mechali M, Blow JJ. Repression of origin assembly in metaphase depends on inhibition of RLF-B/Cdt1 by geminin. Nat Cell Biol. 2001;3:107–13. doi: 10.1038/35055000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi N, Tsutsumi S, Tsuchiya T, Stillman B, Mizushima T. Functions of sensor 1 and sensor 2 regions of Saccharomyces cerevisiae Cdc6p in vivo and in vitro. J Biol Chem. 2002;277:16033–40. doi: 10.1074/jbc.M108615200. [DOI] [PubMed] [Google Scholar]
- Takahashi TS, Wigley DB, Walter JC. Pumps, paradoxes and ploughshares: mechanism of the MCM2–7 DNA helicase. Trends Biochem Sci. 2005;30:437–44. doi: 10.1016/j.tibs.2005.06.007. [DOI] [PubMed] [Google Scholar]
- Takara TJ, Bell SP. Multiple Cdt1 molecules act at each origin to load replication-competent Mcm2–7 helicases. EMBO J. 2011;30:4885–96. doi: 10.1038/emboj.2011.394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takayama Y, Kamimura Y, Okawa M, Muramatsu S, Sugino A, Araki H. GINS, a novel multiprotein complex required for chromosomal DNA replication in budding yeast. Genes Dev. 2003;17:1153–65. doi: 10.1101/gad.1065903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takeda DY, Shibata Y, Parvin JD, Dutta A. Recruitment of ORC or CDC6 to DNA is sufficient to create an artificial origin of replication in mammalian cells. Genes Dev. 2005;19:2827–36. doi: 10.1101/gad.1369805. [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 S, Komeda Y, Umemori T, Kubota Y, Takisawa H, Araki H. Efficient initiation of DNA replication in eukaryotes requires Dpb11/TopBP1-GINS interaction. Mol Cell Biol. 2013;33:2614–22. doi: 10.1128/MCB.00431-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanaka S, Nakato R, Katou Y, Shirahige K, Araki H. Origin association of Sld3, Sld7, and Cdc45 proteins is a key step for determination of origin-firing timing. Curr Biol. 2011a;21:2055–63. doi: 10.1016/j.cub.2011.11.038. [DOI] [PubMed] [Google Scholar]
- Tanaka S, Umemori T, Hirai K, Muramatsu S, Kamimura Y, Araki H. CDK-dependent phosphorylation of Sld2 and Sld3 initiates DNA replication in budding yeast. Nature. 2007;445:328–32. doi: 10.1038/nature05465. [DOI] [PubMed] [Google Scholar]
- Tanaka T, Umemori T, Endo S, Muramatsu S, Kanemaki M, Kamimura Y, Obuse C, Araki H. Sld7, an Sld3-associated protein required for efficient chromosomal DNA replication in budding yeast. EMBO J. 2011b;30:2019–30. doi: 10.1038/emboj.2011.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tardat M, Brustel J, Kirsh O, Lefevbre C, Callanan M, Sardet C, Julien E. The histone H4 Lys 20 methyltransferase PR-Set7 regulates replication origins in mammalian cells. Nat Cell Biol. 2010;12:1086–93. doi: 10.1038/ncb2113. [DOI] [PubMed] [Google Scholar]
- Tatsumi Y, Ezura K, Yoshida K, Yugawa T, Narisawa-Saito M, Kiyono T, Ohta S, Obuse C, Fujita M. Involvement of human ORC and TRF2 in pre-replication complex assembly at telomeres. Genes Cells. 2008;13:1045–59. doi: 10.1111/j.1365-2443.2008.01224.x. [DOI] [PubMed] [Google Scholar]
- Tatsumi Y, Ohta S, Kimura H, Tsurimoto T, Obuse C. The ORC1 cycle in human cells: I. cell cycle-regulated oscillation of human ORC1. J Biol Chem. 2003;278:41528–34. doi: 10.1074/jbc.M307534200. [DOI] [PubMed] [Google Scholar]
- Teer JK, Dutta A. Human Cdt1 lacking the evolutionarily conserved region that interacts with MCM2–7 is capable of inducing re-replication. J Biol Chem. 2008;283:6817–25. doi: 10.1074/jbc.M708767200. [DOI] [PubMed] [Google Scholar]
- Theis JF, Newlon CS. Domain B of ARS307 contains two functional elements and contributes to chromosomal replication origin function. Mol Cell Biol. 1994;14:7652–9. doi: 10.1128/mcb.14.11.7652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thoma F, Bergman LW, Simpson RT. Nuclease digestion of circular TRP1ARS1 chromatin reveals positioned nucleosomes separated by nuclease-sensitive regions. J Mol Biol. 1984;177:715–33. doi: 10.1016/0022-2836(84)90046-9. [DOI] [PubMed] [Google Scholar]
- Thomae AW, Pich D, Brocher J, Spindler MP, Berens C, Hock R, Hammerschmidt W, Schepers A. Interaction between HMGA1a and the origin recognition complex creates site-specific replication origins. Proc Natl Acad Sci U S A. 2008;105:1692–7. doi: 10.1073/pnas.0707260105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thomer M, May NR, Aggarwal BD, Kwok G, Calvi BR. Drosophila double-parked is sufficient to induce re-replication during development and is regulated by cyclin E/CDK2. Development. 2004;131:4807–18. doi: 10.1242/dev.01348. [DOI] [PubMed] [Google Scholar]
- Thommes P, Kubota Y, Takisawa H, Blow JJ. The RLF-M component of the replication licensing system forms complexes containing all six MCM/P1 polypeptides. EMBO J. 1997;16:3312–9. doi: 10.1093/emboj/16.11.3312. [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;161:513–25. doi: 10.1016/j.cell.2015.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Todorov IT, Attaran A, Kearsey SE. BM28, a human member of the MCM2–3–5 family, is displaced from chromatin during DNA replication. J Cell Biol. 1995;129:1433–45. doi: 10.1083/jcb.129.6.1433. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Treisman JE, Follette PJ, O’farrell PH, Rubin GM. Cell proliferation and DNA replication defects in a Drosophila MCM2 mutant. Genes Dev. 1995;9:1709–15. doi: 10.1101/gad.9.14.1709. [DOI] [PubMed] [Google Scholar]
- Triolo T, Sternglanz R. Role of interactions between the origin recognition complex and SIR1 in transcriptional silencing. Nature. 1996;381:251–3. doi: 10.1038/381251a0. [DOI] [PubMed] [Google Scholar]
- Tsakraklides V, Bell SP. Dynamics of pre-replicative complex assembly. J Biol Chem. 2010;285:9437–43. doi: 10.1074/jbc.M109.072504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsuji T, Ficarro SB, Jiang W. Essential role of phosphorylation of MCM2 by Cdc7/Dbf4 in the initiation of DNA replication in mammalian cells. Mol Biol Cell. 2006;17:4459–72. doi: 10.1091/mbc.E06-03-0241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsunematsu T, Takihara Y, Ishimaru N, Pagano M, Takata T, Kudo Y. Aurora-A controls pre-replicative complex assembly and DNA replication by stabilizing geminin in mitosis. Nat Commun. 2013;4:1885. doi: 10.1038/ncomms2859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsuyama T, Tada S, Watanabe S, Seki M, Enomoto T. Licensing for DNA replication requires a strict sequential assembly of Cdc6 and Cdt1 onto chromatin in Xenopus egg extracts. Nucleic Acids Res. 2005;33:765–75. doi: 10.1093/nar/gki226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tugal T, Zou-Yang XH, Gavin K, Pappin D, Canas B, Kobayashi R, Hunt T, Stillman B. The Orc4p and Orc5p subunits of the Xenopus and human origin recognition complex are related to Orc1p and Cdc6p. J Biol Chem. 1998;273:32421–9. doi: 10.1074/jbc.273.49.32421. [DOI] [PubMed] [Google Scholar]
- Valton AL, Hassan-Zadeh V, Lema I, Boggetto N, Alberti P, Saintome C, Riou JF, Prioleau MN. G4 motifs affect origin positioning and efficiency in two vertebrate replicators. EMBO J. 2014;33:732–46. doi: 10.1002/embj.201387506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Deursen F, Sengupta S, De Piccoli G, Sanchez-Diaz A, Labib K. Mcm10 associates with the loaded DNA helicase at replication origins and defines a novel step in its activation. EMBO J. 2012;31:2195–206. doi: 10.1038/emboj.2012.69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vas A, Mok W, Leatherwood J. Control of DNA rereplication via Cdc2 phosphorylation sites in the origin recognition complex. Mol Cell Biol. 2001;21:5767–77. doi: 10.1128/MCB.21.17.5767-5777.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vashee S, Cvetic C, Lu W, Simancek P, Kelly TJ, Walter JC. Sequence-independent DNA binding and replication initiation by the human origin recognition complex. Genes Dev. 2003;17:1894–908. doi: 10.1101/gad.1084203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vashee S, Simancek P, Challberg MD, Kelly TJ. Assembly of the human origin recognition complex. J Biol Chem. 2001;276:26666–73. doi: 10.1074/jbc.M102493200. [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;63:385–96. doi: 10.1016/j.molcel.2016.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vogelauer M, Rubbi L, Lucas I, Brewer BJ, Grunstein M. Histone acetylation regulates the time of replication origin firing. Mol Cell. 2002;10:1223–33. doi: 10.1016/s1097-2765(02)00702-5. [DOI] [PubMed] [Google Scholar]
- Waga S, Zembutsu A. Dynamics of DNA binding of replication initiation proteins during de novo formation of pre-replicative complexes in Xenopus egg extracts. J Biol Chem. 2006;281:10926–34. doi: 10.1074/jbc.M600299200. [DOI] [PubMed] [Google Scholar]
- Walter JC. Evidence for sequential action of cdc7 and cdk2 protein kinases during initiation of DNA replication in Xenopus egg extracts. J Biol Chem. 2000;275:39773–8. doi: 10.1074/jbc.M008107200. [DOI] [PubMed] [Google Scholar]
- Wang B, Feng L, Hu Y, Huang SH, Reynolds CP, Wu L, Jong AY. The essential role of Saccharomyces cerevisiae CDC6 nucleotide-binding site in cell growth, DNA synthesis, and Orc1 association. J Biol Chem. 1999;274:8291–8. doi: 10.1074/jbc.274.12.8291. [DOI] [PubMed] [Google Scholar]
- Wei L, Zhao X. A new MCM modification cycle regulates DNA replication initiation. Nat Struct Mol Biol. 2016;23:209–16. doi: 10.1038/nsmb.3173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei Z, Liu C, Wu X, Xu N, Zhou B, Liang C, Zhu G. Characterization and structure determination of the Cdt1 binding domain of human minichromosome maintenance (Mcm) 6. J Biol Chem. 2010;285:12469–73. doi: 10.1074/jbc.C109.094599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weinreich M, Liang C, Stillman B. The Cdc6p nucleotide-binding motif is required for loading mcm proteins onto chromatin. Proc Natl Acad Sci U S A. 1999;96:441–6. doi: 10.1073/pnas.96.2.441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weinreich M, Stillman B. Cdc7p-Dbf4p kinase binds to chromatin during S phase and is regulated by both the APC and the RAD53 checkpoint pathway. EMBO J. 1999;18:5334–46. doi: 10.1093/emboj/18.19.5334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whittaker AJ, Royzman I, Orr-Weaver TL. Drosophila double parked: a conserved, essential replication protein that colocalizes with the origin recognition complex and links DNA replication with mitosis and the down-regulation of S phase transcripts. Genes Dev. 2000;14:1765–76. [PMC free article] [PubMed] [Google Scholar]
- Wiedemann C, Szambowska A, Hafner S, Ohlenschlager O, Guhrs KH, Gorlach M. Structure and regulatory role of the C-terminal winged helix domain of the archaeal minichromosome maintenance complex. Nucleic Acids Res. 2015;43:2958–67. doi: 10.1093/nar/gkv120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilmes GM, Archambault V, Austin RJ, Jacobson MD, Bell SP, Cross FR. Interaction of the S-phase cyclin Clb5 with an “RXL” docking sequence in the initiator protein Orc6 provides an origin-localized replication control switch. Genes Dev. 2004;18:981–91. doi: 10.1101/gad.1202304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wohlschlegel JA, Dhar SK, Prokhorova TA, Dutta A, Walter JC. Xenopus Mcm10 binds to origins of DNA replication after Mcm2–7 and stimulates origin binding of Cdc45. Mol Cell. 2002;9:233–40. doi: 10.1016/s1097-2765(02)00456-2. [DOI] [PubMed] [Google Scholar]
- Wohlschlegel JA, Dwyer BT, Dhar SK, Cvetic C, Walter JC, Dutta A. Inhibition of eukaryotic DNA replication by geminin binding to Cdt1. Science. 2000;290:2309–12. doi: 10.1126/science.290.5500.2309. [DOI] [PubMed] [Google Scholar]
- Wong HM, Huppert JL. Stable G-quadruplexes are found outside nucleosome-bound regions. Mol Biosyst. 2009;5:1713–9. doi: 10.1039/b905848f. [DOI] [PubMed] [Google Scholar]
- Wu M, Lu W, Santos RE, Frattini MG, Kelly TJ. Geminin inhibits a late step in the formation of human pre-replicative complexes. J Biol Chem. 2014a;289:30810–21. doi: 10.1074/jbc.M114.552935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu R, Wang J, Liang C. Cdt1p, through its interaction with Mcm6p, is required for the formation, nuclear accumulation and chromatin loading of the MCM complex. J Cell Sci. 2012;125:209–19. doi: 10.1242/jcs.094169. [DOI] [PubMed] [Google Scholar]
- Wu Z, Liu J, Yang H, Xiang H. DNA replication origins in archaea. Front Microbiol. 2014b;5:179. doi: 10.3389/fmicb.2014.00179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wuarin J, Buck V, Nurse P, Millar JB. Stable association of mitotic cyclin B/Cdc2 to replication origins prevents endoreduplication. Cell. 2002;111:419–31. doi: 10.1016/s0092-8674(02)01042-5. [DOI] [PubMed] [Google Scholar]
- Wyrick JJ, Aparicio JG, Chen T, Barnett JD, Jennings EG, Young RA, Bell SP, Aparicio OM. Genome-wide distribution of ORC and MCM proteins in S. cerevisiae: high-resolution mapping of replication origins. Science. 2001;294:2357–60. doi: 10.1126/science.1066101. [DOI] [PubMed] [Google Scholar]
- Xu J, Yanagisawa Y, Tsankov AM, Hart C, Aoki K, Kommajosyula N, Steinmann KE, Bochicchio J, Russ C, Regev A, Rando OJ, Nusbaum C, Niki H, Milos P, Weng Z, Rhind N. Genome-wide identification and characterization of replication origins by deep sequencing. Genome Biol. 2012;13:R27. doi: 10.1186/gb-2012-13-4-r27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu W, Aparicio JG, Aparicio OM, Tavare S. Genome-wide mapping of ORC and Mcm2p binding sites on tiling arrays and identification of essential ARS consensus sequences in S. cerevisiae. BMC Genomics. 2006;7:276. doi: 10.1186/1471-2164-7-276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu Y, Gristwood T, Hodgson B, Trinidad JC, Albers SV, Bell SD. Archaeal orthologs of Cdc45 and GINS form a stable complex that stimulates the helicase activity of MCM. Proc Natl Acad Sci U S A. 2016 doi: 10.1073/pnas.1613825113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yabuuchi H, Yamada Y, Uchida T, Sunathvanichkul T, Nakagawa T, Masukata H. Ordered assembly of Sld3, GINS and Cdc45 is distinctly regulated by DDK and CDK for activation of replication origins. EMBO J. 2006;25:4663–74. doi: 10.1038/sj.emboj.7601347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan H, Gibson S, Tye BK. Mcm2 and Mcm3, two proteins important for ARS activity, are related in structure and function. Genes Dev. 1991;5:944–57. doi: 10.1101/gad.5.6.944. [DOI] [PubMed] [Google Scholar]
- Yanagi K, Mizuno T, You Z, Hanaoka F. Mouse geminin inhibits not only Cdt1-MCM6 interactions but also a novel intrinsic Cdt1 DNA binding activity. J Biol Chem. 2002;277:40871–80. doi: 10.1074/jbc.M206202200. [DOI] [PubMed] [Google Scholar]
- Yardimci H, Loveland AB, Habuchi S, Van Oijen AM, Walter JC. Uncoupling of sister replisomes during eukaryotic DNA replication. Mol Cell. 2010;40:834–40. doi: 10.1016/j.molcel.2010.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yeeles JT, Deegan TD, Janska A, Early A, Diffley JF. Regulated eukaryotic DNA replication origin firing with purified proteins. Nature. 2015;519:431–5. doi: 10.1038/nature14285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ying CY, Gautier J. The ATPase activity of MCM2–7 is dispensable for pre-RC assembly but is required for DNA unwinding. EMBO J. 2005;24:4334–44. doi: 10.1038/sj.emboj.7600892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoshida K, Takisawa H, Kubota Y. Intrinsic nuclear import activity of geminin is essential to prevent re-initiation of DNA replication in Xenopus eggs. Genes Cells. 2005;10:63–73. doi: 10.1111/j.1365-2443.2005.00815.x. [DOI] [PubMed] [Google Scholar]
- You Z, Ishimi Y, Mizuno T, Sugasawa K, Hanaoka F, Masai H. Thymine-rich single-stranded DNA activates Mcm4/6/7 helicase on Y-fork and bubble-like substrates. EMBO J. 2003;22:6148–60. doi: 10.1093/emboj/cdg576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- You Z, Masai H. Cdt1 forms a complex with the minichromosome maintenance protein (MCM) and activates its helicase activity. J Biol Chem. 2008;283:24469–77. doi: 10.1074/jbc.M803212200. [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;23:217–24. doi: 10.1038/nsmb.3170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zegerman P, Diffley JF. Phosphorylation of Sld2 and Sld3 by cyclin-dependent kinases promotes DNA replication in budding yeast. Nature. 2007;445:281–5. doi: 10.1038/nature05432. [DOI] [PubMed] [Google Scholar]
- Zellner E, Herrmann T, Schulz C, Grummt F. Site-specific interaction of the murine pre-replicative complex with origin DNA: assembly and disassembly during cell cycle transit and differentiation. Nucleic Acids Res. 2007;35:6701–13. doi: 10.1093/nar/gkm555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang J, Yu L, Wu X, Zou L, Sou KK, Wei Z, Cheng X, Zhu G, Liang C. The interacting domains of hCdt1 and hMcm6 involved in the chromatin loading of the MCM complex in human cells. Cell Cycle. 2010;9:4848–57. doi: 10.4161/cc.9.24.14136. [DOI] [PubMed] [Google Scholar]
- Zhang X, Shaw A, Bates PA, Newman RH, Gowen B, Orlova E, Gorman MA, Kondo H, Dokurno P, Lally J, Leonard G, Meyer H, Van Heel M, Freemont PS. Structure of the AAA ATPase p97. Mol Cell. 2000;6:1473–84. doi: 10.1016/s1097-2765(00)00143-x. [DOI] [PubMed] [Google Scholar]
- Zhang Z, Hayashi MK, Merkel O, Stillman B, Xu RM. Structure and function of the BAH-containing domain of Orc1p in epigenetic silencing. EMBO J. 2002;21:4600–11. doi: 10.1093/emboj/cdf468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong W, Feng H, Santiago FE, Kipreos ET. CUL-4 ubiquitin ligase maintains genome stability by restraining DNA-replication licensing. Nature. 2003;423:885–9. doi: 10.1038/nature01747. [DOI] [PubMed] [Google Scholar]
- Zhou C, Huang SH, Jong AY. Molecular cloning of Saccharomyces cerevisiae CDC6 gene. Isolation, identification, and sequence analysis. J Biol Chem. 1989;264:9022–9. [PubMed] [Google Scholar]
- Zou L, Mitchell J, Stillman B. CDC45, a novel yeast gene that functions with the origin recognition complex and Mcm proteins in initiation of DNA replication. Mol Cell Biol. 1997;17:553–63. doi: 10.1128/mcb.17.2.553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zwerschke W, Rottjakob HW, Kuntzel H. The Saccharomyces cerevisiae CDC6 gene is transcribed at late mitosis and encodes a ATP/GTPase controlling S phase initiation. J Biol Chem. 1994;269:23351–6. [PubMed] [Google Scholar]