40
DocuServe Electronic Delivery Cover Sheet WARNING CONCERNING COPYRIGHT RESTRICTIONS The copyright law of the United States (Title 17, United States Code) governs the making of photocopies or other reproductions of copyrighted materials. Under certain conditions specified in the law, libraries and archives are authorized to furnish a photocopy or other reproduction. One of these specified conditions is that the photocopy or reproduction is not to be "used for any purpose other than private study, scholarship, or research". If a user makes a request for, or later uses, a photocopy or reproduction for purposes in excess of "fair use", that user may be liable for copyright infringement. This institution reserves the right to refuse to accept a copying order if, in its judgment, fulfillment of the order would involve violation of copyright law. Caltech Library Services

Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

 

DocuServe  

Electronic Delivery Cover Sheet  

 

 

WARNING CONCERNING COPYRIGHT RESTRICTIONS 

 

The copyright law of the United States (Title 17, United States Code) governs  the  making  of  photocopies  or  other  reproductions  of copyrighted materials.  Under certain conditions specified in the law, libraries and archives are authorized to furnish a photocopy or other reproduction.    One  of  these  specified  conditions  is  that  the photocopy or reproduction  is not to be "used for any purpose other than  private  study,  scholarship,  or  research".    If  a  user  makes  a request for, or  later uses, a photocopy or reproduction for purposes in  excess  of  "fair  use",  that  user  may  be  liable  for  copyright infringement.  This institution reserves the right to refuse to accept a copying  order  if,  in  its  judgment,  fulfillment  of  the  order  would involve violation of copyright law. 

 

Caltech Library Services 

Page 2: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Full Terms & Conditions of access and use can be found athttp://www.tandfonline.com/action/journalInformation?journalCode=ibmg20

Critical Reviews in Biochemistry and Molecular Biology

ISSN: 1040-9238 (Print) 1549-7798 (Online) Journal homepage: http://www.tandfonline.com/loi/ibmg20

Mechanisms and regulation of DNA replicationinitiation in eukaryotes

Matthew W. Parker, Michael R. Botchan & James M. Berger

To cite this article: Matthew W. Parker, Michael R. Botchan & James M. Berger (2017)Mechanisms and regulation of DNA replication initiation in eukaryotes, Critical Reviews inBiochemistry and Molecular Biology, 52:2, 107-144, DOI: 10.1080/10409238.2016.1274717

To link to this article: https://doi.org/10.1080/10409238.2016.1274717

Published online: 17 Jan 2017.

Submit your article to this journal

Article views: 892

View related articles

View Crossmark data

Citing articles: 8 View citing articles

Page 3: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

REVIEW ARTICLE

Mechanisms and regulation of DNA replication initiation in eukaryotes

Matthew W. Parkera, Michael R. Botchanb and James M. Bergera

aDepartment of Biophysics and Biophysical Chemistry, Johns Hopkins University School of Medicine, Baltimore, MD, USA; bDepartmentof Molecular and Cell Biology, University of California Berkeley, Berkeley, CA, USA

ABSTRACTCellular DNA replication is initiated through the action of multiprotein complexes that recognizereplication start sites in the chromosome (termed origins) and facilitate duplex DNA meltingwithin these regions. In a typical cell cycle, initiation occurs only once per origin and each roundof 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 ofthe helicase by its incorporation into a complex known as the CMG. Recent work has begun toreveal the details of an orchestrated and sequential exchange of initiation factors on DNA thatgive rise to a replication-competent complex, the replisome. Here, we review the molecularmechanisms that underpin eukaryotic DNA replication initiation – from selecting replication startsites to replicative helicase loading and activation – and describe how these events are often dis-tinctly regulated across different eukaryotic model organisms.

ARTICLE HISTORYReceived 17 October 2016Revised 14 December 2016Accepted 16 December 2016

KEYWORDSDNA replication; ORC; Cdc6;initiator; Cdt1; MCM2-7;CMG; helicase

Introduction

The success of biological organisms depends on thefaithful transmission of genetic information from parentto progeny. All life-forms store their genetic content inthe form of nucleic acids, and the replication and dis-semination of this information forms the fundamentalbasis of inheritance. In cells, the process of replicationinvolves two primary tasks: (1) the separation of duplexDNA into two single-stranded templates and (2) semi-conservative replication of each strand. These eventsare coupled with cell division to produce progenywith essentially identical copies of the parent’s geneticinformation. Through all cellular lineages, a conserveddivision of labor has been applied to the process ofDNA replication, such that separable tasks (e.g. startsite selection, duplex unwinding, DNA synthesis) areallocated to different, albeit sometimes overlapping,factors. Although this basic framework is conservedthroughout the bacterial, archaeal, and eukaryotickingdoms, there has been significant evolutionary diver-sification of the molecules that complete each task, tothe point where it is now clear that aspects of the repli-cative machinery emerged twice, independently, duringcellular evolution (Edgell & Doolittle, 1997; Leipe et al.,1999). Replication initiation in eukaryl species has

become particularly elaborated by disparate forms ofregulation to meet the specific demands of multicellu-larity, development and large genome size.

Replication is started by a trans-acting “initiator” fac-tor that directs, in both space and time, loading of thereplicative machinery onto particular genomic lociknown as origins. In general, the number of originsscales with genome size, thereby ensuring that chromo-some duplication can be carried out on a physiologic-ally manageable timescale (Gilbert, 2004). Bacteria, aswell as certain archaea, frequently initiate replicationwith a single chromosomal start site (Costa et al., 2013;Wu et al., 2014b). Conversely, some archaeal chromo-somes possess multiple origins (Wu et al., 2014b), as doall eukaryotic genomes (the 12 Mbp Saccharomycescerevisiae genome contains around 400 origins and the3 Gbp human genome �30,000–50,000 origins(Leonard & Mechali, 2013)). The large number of originsand the need to coordinate initiation across these sitesrepresents a fundamental challenge to DNA replicationin eukaryotes; other factors, such as the use of multiplelinear chromosomes, as opposed to a single circularchromosome, add additional replicative complexity.Replication is particularly problematic in multicellulareukaryotes, where the process of development can alter

CONTACT James M. Berger [email protected] Department of Biophysics and Biophysical Chemistry, Johns Hopkins University School of Medicine,Baltimore, MD, USA; Michael R. Botchan [email protected] Department of Molecular and Cell Biology, University of California Berkeley,Berkeley, CA, USA� 2017 Informa UK Limited, trading as Taylor & Francis Group

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY, 2017VOL. 52, NO. 2, 107–144http://dx.doi.org/10.1080/10409238.2016.1274717

Page 4: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

replication timing and frequency, and in the context ofcellular differentiation, which changes the chromatinlandscape and requires the coordinated transmission ofepigenetic marks. Despite these and other challenges,more than 15 trillion repetitive rounds of DNA replica-tion and cell division are successfully executed on thedevelopmental path of a fertilized human embryo tothe adult human body (Bianconi et al., 2013). Althoughour understanding of the replicative process is far fromcomplete, we are beginning to understand how eukar-yotes utilize a variety of sequential and redundant regu-latory mechanisms to achieve this biological feat.

The eukaryotic replisome is built from the regulatedand stepwise assembly of multiple intermediary replica-tion factor complexes. In S. cerevisiae, 42 individual pro-teins are sufficient to fully reconstitute DNA replicationin vitro, and since many of these proteins functionwithin large macromolecular assemblies, fewer than 15pre-assembled replication factors are required (Yeeleset al., 2015). In short, replication initiation entails foursteps (Figure 1): (1) demarcation of start sites by theOrigin Recognition Complex (ORC) and the Cdc6 heli-case-loader; (2) reiterative loading of an inactive form ofthe replicative helicase, MCM2–7, by ORC·Cdc6 and theCdt1 chaperone to form the pre-replication complex(pre-RC); (3) helicase activation by the formation of theCdc45·MCM2–7·GINS (CMG) complex (the pre-initiationcomplex, pre-IC) and (4) generation of a bidirectionalreplication fork that depends on prior origin melting by

the MCM2–7 complex and on the tethering of DNA pol-ymerases and additional accessory factors to the repli-cative helicase. Here, we review the molecularmechanisms underpinning eukaryotic replication initi-ation, from origin specification to helicase activation.This review, although focused on eukaryotic mecha-nisms, will, as needed, refer studies of the archaeal sys-tem to fill critical gaps in knowledge. Given theextensive number of publications in the field, we apolo-gize to those colleagues whose work is not referenceddue to space limitations.

Origins of replication

Origins of replication are chromosomal regions thatrecruit replication initiators for facilitating assembly ofthe replication machinery (Francois Jacob, 1963). Thedefining features of eukaryotic origins are complicatedand continuously evolving; for more thorough coveragewe refer the reader to a number of excellent reviews onthe topic (Creager et al., 2015; Leonard & Mechali, 2013;MacAlpine & Bell, 2005). Here, we briefly discuss themost salient features of origins in S. cerevisiae, S. pombeand metazoans, with a particular focus on details pertin-ent to later topics of discussion.

S. cerevisiae origins of replication

Saccharomyces cerevisiae origins of replication were ini-tially identified as chromosomal regions capable of

Figure 1. Mechanistic outline of DNA replication initiation in eukarya. During the G1 phase of the cell cycle, an origin-boundORC·Cdc6 complex together with Cdt1 facilitates the sequential recruitment and loading of two MCM2–7 complexes into a stabledouble hexamer that encircles duplex DNA (pre-RC). At the onset of S phase, the helicase is activated, leading to origin unwind-ing. The recruitment of other initiation factors (Cdc45 and GINS, the pre-IC) and double-hexamer dissolution activate the helicaseto drive fork progression as a single-stranded DNA-bound Cdc45·MCM2–7·GINS (CMG) complex. A color version of this figure isavailable at www.tandfonline.com/ibmg.

108 M. W. PARKER ET AL.

Page 5: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

conferring replicative properties to exogenous plas-mids (Stinchcomb et al., 1979). These AT-rich, autono-mous replication sequences (ARSs) function asreplication start sites (Brewer & Fangman, 1987;Huberman et al., 1987,1988) and serve to recruitthe eukaryotic initiator, ORC (Bell & Stillman, 1992).The ARS contains a number of essential elements, themost important of which is the 11-basepair “A ele-ment”, which constitutes the ARS consensus sequence(ACS) and represents the primary site of initiator bind-ing (Marahrens & Stillman, 1992; Rao et al., 1994; Rao& Stillman, 1995; Theis & Newlon, 1994). Notably,among the different model eukaryotic systems usedfor studying replication, only S. cerevisiae appears toutilize a specific consensus sequence (Figure 2).

Although there are over 12,000 ACS-like sequencesin the yeast genome, only about 400 facilitate replica-tion initiation (Nieduszynski et al., 2006; Wyrick et al.,

2001; Xu et al., 2006). This low usage of possibleS. cerevisiae origins (<5%) derives in part from an add-itional level of origin specification that is imposed bythe local chromatin structure. Saccharomyces cerevisiaeorigins, like those of other eukaryotes, are maintainedas nucleosome-free regions (NFRs) (Berbenetz et al.,2010; Eaton et al., 2010). However, surrounding an ARS,ORC-dependent nucleosome phasing directly affectsthe efficiency of origin usage (Berbenetz et al., 2010;Lipford & Bell, 2001; Simpson, 1990; Thoma et al., 1984),and the eviction of ORC results in nucleosomeencroachment into ARS regions (Eaton et al., 2010;Thoma et al., 1984). Mechanistically, NFRs allow ORCaccess to DNA, while phased nucleosomes provide add-itional favorable sites for binding ORC (Hizume et al.,2013; Muller et al., 2010). Thus, S. cerevisiae origins aredefined by ORC binding to both an ACS and specificchromatin features (Hoggard et al., 2013; Leonard &Mechali, 2013).

S. pombe origins of replication

Genome-wide studies demonstrate that the S. pombegenome contains around 400 origin sequences that aregenerally nucleosome-free (Givens et al., 2012; Xu et al.,2012), AT-rich, and around 1 kilobase long (Dai et al.,2005; Heichinger et al., 2006; Segurado et al., 2003).These findings are consistent with biochemical studiesshowing that origin usage in fission yeast depends onclustered stretches of adenine and thymine (Clyne &Kelly, 1995; Dai et al., 2005; Kim et al., 2001; Okunoet al., 1999). Origin selection in S. pombe is facilitated bya species-specific insertion in the Orc4 subunit of ORCthat encodes a DNA-binding element that specificallyrecognizes the minor groove of AT-rich sequences(Figure 2) (Chuang & Kelly, 1999; Moon et al., 1999) .

Metazoan origins of replication

Unlike budding yeast ORC, which shows a degree ofsequence-specificity, metazoan ORC binds DNA promis-cuously (Remus et al., 2004; Vashee et al., 2003). Thisbehavior is consistent with the observation that meta-zoan replication initiates from diverse sequences(Heinzel et al., 1991; Hyrien & Mechali, 1993; Mechali &Kearsey, 1984). Despite origin sequence variability, thegenome-wide analysis of replication start sites hasrevealed some common patterns in metazoan origins.As with budding and fission yeast, metazoan ORC bindsto NFRs in the genome (Eaton et al., 2011; Karnani et al.,2010; MacAlpine et al., 2010), which in turn favorablycontributes to assembly of the replication machinery(Lubelsky et al., 2011). Interestingly, G-rich sequences

Figure 2. Molecular details of eukaryotic origins and mecha-nisms of ORC binding. (gray) Saccharomyces cerevisiae originsare distinctive among eukaryotes for conforming to a consen-sus sequence, the ACS. ScORC can bind the ACS directly andspecifically, although interactions between the Orc1-BAHdomain and nucleosomes can also modulate ORC origin selec-tion. (PURPLE) Although they do not possess a strict consen-sus sequence, S. pombe origins are AT-rich. SpORC specificallybinds such sites using a domain insertion unique to SpOrc4that encodes a DNA-binding AT-hook motif. (green) MetazoanORC can be targeted to chromosomes through a variety ofmechanisms, including the Orc1 BAH domain, the DNA-bind-ing TFIIB domain of Orc6, and through interactions with chro-matin-associated factors. A majority of metazoan origins arealso predicted to contain G-quadruplex secondary structureelements, but how this feature affects ORC binding is cur-rently unclear. A color version of this figure is available atwww.tandfonline.com/ibmg.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 109

Page 6: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

and CpG islands are highly enriched in metazoan ori-gins (Cadoret et al., 2008; Delgado et al., 1998; Prioleau,2009; Sequeira-Mendes et al., 2009) and have been pro-posed to serve two purposes: NFR maintenance(Huppert & Balasubramanian, 2007; Fenouil et al., 2012;Wong & Huppert, 2009) and the favoring of G-quadra-plex formation (Cayrou et al., 2011, 2012, 2015; Valtonet al., 2014). Preliminary analyses suggest that ORC maypreferentially associate with G-rich elements (Hoshinaet al., 2013; Zellner et al., 2007), indicating that a con-served structural feature in DNA, rather than a specificconsensus sequence, may aid ORC binding in metazo-ans. As in S. cerevisiae, metazoan origin selection byORC is further fine-tuned by direct interactions withnucleosomes and chromatin-associated factors. Forexample, the N-terminal bromo-adjacent homology(BAH) domain in Orc1 directly interacts with histones todirect origin usage (Figure 2) (Kuo et al., 2012; Mulleret al., 2010; Noguchi et al., 2006).

Although ORC can be targeted to specific genomicloci, origins cannot be strictly defined by the position ofORC binding on chromatin. Indeed, once loaded ontoDNA by ORC, the MCM2–7 helicase (which eventuallynucleates replisome assembly following duplex melting)appears to lack positional restraints and is free to eitherdiffuse away from ORC (Evrin et al., 2009; Remus et al.,2009) or be forcibly displaced by other chromatin-local-ized cellular processes (such as the transcriptionalmachinery) (Edwards et al., 2002; Gros et al., 2015;Powell et al., 2015; Ritzi et al., 1998). Thus, origins ineukaryotes must be defined flexibly, as the site of initi-ator binding does not always reflect the site of replica-tion initiation.

Overall, both yeast and metazoa contain conservedsequence elements at origins that are known or pro-posed to guide ORC binding. In yeast, these elementsare encoded within the DNA primary sequence and inmetazoa potentially by a propensity to form distinctivesecondary structures. Chromatin context plays an add-itional critical, but poorly understood role in originusage in all eukaryotes. Understanding how these cis-and trans-acting origin elements interface is an import-ant and active area of future research.

The origin recognition complex (ORC)

Eukaryotic origins direct the recruitment of the originrecognition complex (ORC), a conserved heterohexa-meric protein assembly identified for its ability to specif-ically recognize the double-stranded form of the yeastACS (Bell & Stillman, 1992). Upon recruitment tochromosomal replication start sites, ORC binds an add-itional factor, Cdc6, as a necessary prerequisite to

helicase loading. Despite its centrality to ORC function,the mechanism of DNA binding by the initiator haslong remained ambiguous. However, structural studiesfrom archaea and eukaryotes have revealed a conservedmechanism for the association of ORC with DNA thatinforms not only our understanding of how theORC·Cdc6 initiator stably binds replication start sites,but also how ORC mediates downstream helicase load-ing events (Bleichert et al., 2015; Dueber et al., 2007;Gaudier et al., 2007; Sun et al., 2013).

The origin recognition complex and Cdc6

ORC was first identified by fractionation of ARS-bindingproteins in budding yeast (Bell & Stillman, 1992).Although many other ARS-binding factors had beenidentified previously (Buchman et al., 1988; Diffley &Stillman, 1988; Jazwinski & Edelman, 1982; Shore et al.,1987; Sweder et al., 1988), ORC proved uniquely able tobind the ACS (Diffley & Cocker, 1992; Li & Herskowitz,1993; Marahrens & Stillman, 1992,), and temperaturesensitive mutants exhibited cell-cycle arrest at a stageconsistent with a role in the early aspects of DNA repli-cation (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 demon-strated with the identification of orthologs in S. pombe(Grallert & Nurse, 1996; Muzi-Falconi & Kelly, 1995),Drosophila melanogaster (Gossen et al., 1995; Landiset al., 1997), Xenopus laevis (Carpenter et al., 1996;Romanowski et al., 1996b) and humans (Gavin et al.,1995). The mechanism of ORC function has now beeninvestigated across multiple model organisms, revealingthat the core components, subunit organization andfunction of ORC are broadly conserved (although notuniversally, particularly in protozoa (El-Sayed et al.,2005)). Despite this conservation, there do exist certainspecies-specific alterations that appear to have gener-ated notable functional differences.

Saccharomyces cerevisiae ORC is a roughly 400 kDaassembly composed of six proteins (Orc1–6) named indescending order of molecular weight (Bell et al., 1993).Orc1–5 exhibit relatively good conservation across spe-cies and are members of the ATPases associated withdiverse cellular activities (AAAþ) superfamily of pro-teins; Orc1–5 also contain a C-terminal winged-helix(WH) domain (Bell et al., 1995; Liu et al., 2000; Loo et al.,1995; Muzi-Falconi & Kelly, 1995). Overall, the Orc1–5AAAþ and WH domains, which account for a bit overhalf the total mass of ORC, display an average of�45–50% similarity and 25–30% identity between yeastand human orthologs (Speck et al., 2005; Tugal et al.,1998). A majority of AAAþ family members bind and

110 M. W. PARKER ET AL.

Page 7: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

hydrolyze ATP through the formation of a composite,bipartite active site that is generated between neigh-boring AAAþ protomers upon subunit oligomerization(Guenther et al., 1997; Lenzen et al., 1998; Putnam et al.,2001). AAAþ proteins, like the broad P-loop family ofNTPases to which they belong (Iyer et al., 2004), containthe so-called “Walker A” and “Walker B” signaturesequence motifs that contribute to nucleotide bindingand hydrolysis, respectively. A third motif, generally crit-ical to AAAþ ATPase function, is the arginine finger,which is presented in trans to the neighboring subunitto reconstitute an active ATPase site (Neuwald et al.,1999; Zhang et al., 2000). While many of the structuralelements important for ATP binding are conserved inOrc1, Orc4 and Orc5, only Orc1 has been found to pos-sess ATPase activity (Bleichert et al., 2015; Klemm et al.,1997; Makise et al., 2003; Ranjan & Gossen, 2006;Siddiqui & Stillman, 2007); Orc2 and Orc3 retain degen-erate AAAþ scaffolds that lack functional active sitemotifs altogether (Bleichert et al., 2015; Clarey et al.,2006; Speck et al., 2005). Intriguingly, although manyAAAþ family members function as toroidal hexamericassemblies (reviewed in (Hanson & Whiteheart, 2005)),ORC retains only five proteins with AAAþ domains.Orc6 lacks a AAAþ domain (Balasov et al., 2007;Bleichert et al., 2013; Chesnokov et al., 2003) and its pri-mary sequence is only weakly conserved between yeastand human, making it the least conserved ORC subunit(Dhar & Dutta, 2000). Nonetheless, certain elements ofOrc6 are conserved across species, including an N-ter-minal TFIIB-like domain and a short conserved region atthe extreme C-terminus of the protein (Bleichert et al.,2013; Liu et al., 2011).

Cdc6 forms a complex with ORC at origins and isrequired for initiator function. Cdc6 was first identifiedin S. cerevisiae mutant screens (Hartwell, 1976) and waslater found to have a role in replication initiation(Palmer et al., 1990), with functional requirements priorto S-phase (Hogan & Koshland, 1992; Kelly et al., 1993;Zwerschke et al., 1994). A genetic interaction betweenORC and Cdc6 suggested a coordinated activity forthese factors (Liang et al., 1995), with supporting datademonstrating an ORC-dependent recruitment of Cdc6to origins that relies on a direct interaction between theproteins (Cocker et al., 1996; Coleman et al., 1996;Grallert & Nurse, 1996; Kong et al., 2003; Leatherwoodet al., 1996; Liang et al., 1995; Santocanale & Diffley,1996). Importantly, analysis of the Cdc6 primarysequence reveals conserved nucleotide binding andhydrolysis motifs (Lisziewicz et al., 1988; Zhou et al.,1989; Zwerschke et al., 1994), as well as close homologyto ORC’s AAAþ subunits, particularly Orc1 (Bell et al.,1995; Quintana et al., 1997). Cdc6 associates with

chromatin-bound ORC in an ATP-dependent manner(Coster et al., 2014; Evrin et al., 2013; Kang et al., 2014;Kneissl et al., 2003; Perkins & Diffley, 1998; Ticau et al.,2015), an interaction that activates Cdc6’s ATPase activ-ity and is consistent with the reconstitution of canonicalAAAþ interactions (Randell et al., 2006; Speck &Stillman, 2007). Thus, Cdc6 recruitment provides a sixthAAAþ subunit to the initiator complex overall.

Structure of ORC and ORC·Cdc6

Orc/Cdc6 homologs have been identified in all eukary-otic and archaeal species analyzed (Aves et al., 2012). Inmany archaea, the initiation factors are geneticallystreamlined such that certain species possess only a sin-gle Orc/Cdc6 gene (Barry & Bell, 2006) and the relativesimplicity of the archaeal system has been exploited tohelp understand the structure and function of eukary-otic Orc homologs. A conserved three-domain architec-ture is observed for Orc and Cdc6 homologs, with twodomains at the N-terminus forming the central AAAþmodule and a third at the C-terminus comprising aloosely-tethered WH domain (Figure 3(A)) (Bleichertet al., 2015; Dueber et al., 2007; Liu et al., 2000; Gaudieret al., 2007; Singleton et al., 2004). The structure of arch-aeal Orc in complex with DNA has shown that both theWH and AAAþ domains contact DNA (Dueber et al.,2007; Gaudier et al., 2007), and biochemical studieshave demonstrated that contacts mediated by theAAAþ domain contribute to origin specificity (Dueberet al., 2011). Surprisingly, structures of Orc/DNA com-plexes have not revealed evidence for the formation ofan ATPase-competent Orc dimer (Dueber et al., 2007;Gaudier et al., 2007), despite the existence of a con-served arginine finger in these proteins, which wouldotherwise suggest that oligomerization might occur.The mechanism that promotes formation of a catalytic-ally active ATPase in the archaeal system is unclear.

Saccharomyces cerevisiae Cdc6 forms a stable com-plex with ORC in the presence of DNA and ATP(Mizushima et al., 2000; Seki & Diffley, 2000; Speck et al.,2005; Randell et al., 2006; Wang et al., 1999). By using anon-hydrolyzable ATP analog, a stable Cdc6·ORC com-plex can also be trapped in the absence of DNA (Specket al., 2005). Low-resolution 3D electron microscopyreconstructions of eukaryotic ORC have revealed anelongated, crescent-shaped particle that, in the pres-ence of Cdc6 and ATPcS, transforms into a closed ringwith a large central cavity (Clarey et al., 2006, 2008;Speck et al., 2005; Sun et al., 2012). Structural investiga-tion of D. melanogaster ORC has demonstrated a sub-unit order of Orc1!Orc4!Orc5!Orc3!Orc2 aroundthe ORC ring (Bleichert et al., 2015), with a physical gap

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 111

Page 8: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

between the terminal Orc subunits that accommodatesCdc6 and is consistent with the direct interactionobserved between Cdc6 and Orc1 (Sun et al., 2012,2013; Wang et al., 1999). The C-terminus of Orc6 hasbeen found to tether this subunit to a conserveddomain insertion within Orc3 (Bleichert et al., 2013,2015); S. cerevisiae Orc6 also binds Orc2 (Chen et al.,2008; Sun et al., 2012), possibly through a region that isspecific to fungal homologs (Bleichert et al., 2013). Theadjoining nature of Orc1 and Orc4 within the ternarycomplex is consistent with the known formation of ajoint ATPase site between the two subunits (Bowerset al., 2004; Chesnokov et al., 2001; Giordano-Coltartet al., 2005; Klemm et al., 1997).

The ORC·Cdc6 assembly represents the functional ini-tiator complex at origins. Interestingly, ORC itself exhib-its differing levels of stability across species, suggestingthat, in certain cases, ORC subcomplexes may besequentially recruited. Indeed, unlike D. melanogasterand S. cerevisiae ORC, which form stable heterohex-amers (Bell & Stillman, 1992; Chesnokov et al., 1999),human and X. laevis ORC appear to have alternativecore subcomplexes. Human Orc1 and Orc6 loosely asso-ciate with an Orc2–5 core (Dhar et al., 2001; Siddiqui &Stillman, 2007; Vashee et al., 2001), whereas in X. laevis,the Orc6 subunit is labile (Gillespie et al., 2001). Theweak interactions of certain Orc subunits likely play an

important role in regulation. For example, vertebrateOrc1 is selectively released from chromatin after initi-ation (Li et al., 2004; Natale et al., 2000; Rowles et al.,1999), which helps to prevent re-initiation and providesa means to alter origin usage in a developmental- ordifferentiation-dependent fashion (Li & DePamphilis,2002). The ability of either Orc1 or Orc6 to dissociatefrom ORC without disrupting the remaining core com-plex is consistent with the terminal position of Orc1within the core ring and the peripheral binding site ofOrc6; one exception to this trend occurs in S. pombe,whereby an ORC pentamer can be purified that lacksOrc4, an internal component of the ORC ring (Kong &DePamphilis, 2001; Moon et al., 1999). Whether S.pombe ORC retains additional stabilizing elements thatcan compensate for the absence of Orc4 in such instan-ces is not known.

A recent atomic-resolution structure of D. mela-nogaster ORC (Bleichert et al., 2015) has helped to clarifyboth our understanding of ORC organization and mech-anistic models for origin engagement and helicase load-ing. ORC adopts a two-tiered, notched ring architecturein which the WH domains of Orc1–5 sit atop a layer ofAAAþ subunits (Figure 3(B)). Interestingly, the arrange-ment of AAAþ and WH domain contacts is domainswapped in ORC, such that the WH domain of Orc1rests on the AAAþ region of Orc4, the WH domain of

Figure 3. ORC architecture. (A) Cdc6/Orc homologs are characterized by three domains, two of which form the AAAþ module(green) and a third that encodes a winged-helix (WH) domain (gray). Bound nucleotide is shown as sticks (PDB ¼1FNN). (B) TheD. melanogaster ORC heterohexamer is a crescent-shaped molecule with the AAAþ (green surface) and WH (gray cartoon)domains forming a domain-swapped arrangement. Orc6 is bound by a domain insertion in the AAAþ domain of Orc3. Althoughthe Orc1/Orc4 active site is required for activity, in the D. melanogaster structure Orc1 is disengaged from Orc4 and positionedabove the plane of the AAAþ ring (PDB ¼4XGC). A color version of this figure is available at www.tandfonline.com/ibmg.

112 M. W. PARKER ET AL.

Page 9: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Orc4 sits on the AAAþ region of Orc5, and so forth. Theopen-ended, pentameric AAAþ core of ORC in principleprovides an opportunity for the formation of fourbipartite AAAþ interfaces. Within the D. melanogasterORC structure, however, only one of the observed sub-unit interfaces evinces a typical AAAþ active-site config-uration (Orc4/Orc5). Two others (Orc5/3 and Orc3/2)approximate the correct protomer alignment but failto reconstitute a canonical active site; in the case ofOrc3/2, the two subunits lack the amino acids necessaryto bind ATP, as predicted. Of the remaining possibleAAAþ interactions, Orc1 was unexpectedly found to berotated more than 90 degrees out-of-plane from Orc4,rendering this catalytic center inoperative (Figure 3(B)).This finding was surprising, as Orc1 alone is capable ofboth binding and hydrolyzing ATP, using an argininefinger donated by Orc4 (Chesnokov et al., 2001; Klemmet al., 1997). Given that the D. melanogaster Orc1 con-formation seen crystallographically is also seen in 3Delectron microscopy reconstructions of ORC (Bleichertet al., 2013, 2015), this observation suggests that meta-zoan ORC can transition between at least two confor-mations, an autoinhibited and active conformation. Inthe future, it will be important to understand how thecell regulates the equilibria of ORC between thesestates, as well as whether this conformational transitionis preserved in other ORC homologs.

A two-state model for ORC origin recognition

Despite the fundamental role of ORC in origin selectionand recognition, the mechanism by which ORC associ-ates with DNA has remained highly enigmatic. The pres-ence of numerous interspecies peculiarities, such asdivergent origin features, species-specific DNA-bindingelements, and the effect of chromatin-bound trans-act-ing factors, have all challenged our understanding ofhow ORC is recruited to and stably binds origins.However, an analysis of ORC behavior across speciessuggests that in all cases, the origin-binding propertiesof ORC can be interpreted within a model containing atleast two states: a transient ORC recruitment event thatis mediated through diverse and sometimes species-specific interactions, and a second, mechanistically con-served step that positions ORC for productive originengagement and that leads to stable Cdc6 association.We will first discuss the conserved mechanism by whichORC stably associates with origins and then detail theinteractions that facilitate ORC recruitment.

Structural studies of origin-bound archaeal Orc haverevealed a coordinated role for both the AAAþ and WHdomains in DNA binding (Dueber et al., 2007; Gaudieret al., 2007). The WH domains show extensive contacts

with origin DNA using a canonical helix-turn-helix (HTH)and b-hairpin wing interface, and a near compete lossof DNA binding by archaeal Orc is observed upon muta-tion or deletion of this region (Dueber et al., 2011;Singleton et al., 2004). Interestingly, the WH-DNA inter-action positions the AAAþ domain in an orientationwhere a characteristic a-helical insertion within the initi-ator/helicase loader subgroup of AAAþ proteins, theinitiator specific motif (ISM), contacts DNA. Given thedifferent subunit compositions and oligomeric statesobserved between archaeal (monomer) and eukaryoticinitiators (heterohexamer) it was initially unclear towhat extent the mechanism of archaeal origin bindingwould be conserved. However, superposing the struc-ture of DNA-bound archaeal Orc onto the D. mela-nogaster ORC crystal structure reveals nearly perfect co-axial positioning of the DNA within the ORC centralchannel (Bleichert et al., 2015). Notably, the eukaryoticISMs and the wings of the WH domains each form acontiguous DNA-binding surface that lines the ORC cen-tral channel, and are thus positioned to engage DNA ina manner similar to that of archaeal Orc (although notidentically, as the HTH motif of ORC’s WH domains areburied between inter-subunit contacts in the complex(Bleichert et al., 2015)). This modeling also accounts foran unresolved region of density in an early 3D electronmicroscopy reconstruction of an ORC·Cdc6·DNA com-plex (Sun et al., 2012). Overall, these data indicate thatarchaeal and eukaryotic ORC engage DNA by a gener-ally conserved mechanism.

The insights gleaned from the available structuralstudies suggest a conserved mechanism for stable ori-gin association by eukaryotic ORC. In this model, duplexDNA is loaded laterally into the ORC central channelthrough the discontinuity between Orc1 and Orc2(Speck et al., 2005). This gap is then blocked off by thesubsequent recruitment of Cdc6, generating a stable,closed-ring conformation that encircles origin DNA(Bleichert et al., 2015). Domain swapping between theWH domain of Orc2 and the AAAþ region of Cdc6 (andbetween the WH domain of Cdc6 and the AAAþ elem-ent of Orc1) are predicted to form, stabilizing the com-plex (Bleichert et al., 2015). The ability of the ORC·Cdc6complex to encircle DNA would be predicted to under-pin ORC’s observed persistence at origins (Duzdevichet al., 2015; Speck et al., 2005) and is consistent with theobservation that both yeast and metazoan Cdc6 canstabilize ORC on DNA (Harvey & Newport, 2003;Houchens et al., 2008).

Interestingly, an additional DNA-binding element hasbeen identified between the N-terminal BAH and AAAþdomain of S. cerevisiae Orc1 that is essential for ARSbinding (Kawakami et al., 2015). While the related

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 113

Page 10: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

residues are largely conserved in metazoan Orc1, theyare positioned outside of the ORC central pore and thusit is unclear how this interaction is coordinated withthe encirclement mechanism described here, orwhether it contributes only to the initial recruitmentof ORC to DNA.

Diverse mechanisms for origin recruitment of ORC

Although the encirclement model accounts for stableorigin binding, eukaryotic ORC has been observed tointeract with DNA in a mode that exhibits fast kinetics(Duzdevich et al., 2015; Harvey & Newport, 2003;Houchens et al., 2008; Remus et al., 2004) and, in S.cerevisiae, allows for a one-dimensional linear search forbona fide origin sites (Duzdevich et al., 2015). This initialrecruitment of ORC to DNA likely functions as animportant intermediate on the path towards stableorigin binding. Interestingly, substantial mechanisticplasticity appears to have been introduced to therecruitment step, co-evolving in certain cases with spe-cies-specific features.

In S. pombe ORC, the preference for AT-rich originscorrelates with a unique domain insertion in Orc4 com-prising nine AT-hook motifs (Chuang & Kelly, 1999;Moon et al., 1999), a DNA-binding element that facili-tates interactions with the minor groove of AT-rich DNAsequences (Aravind & Landsman, 1998). Notably, theAT-hook motif is absent in S. cerevisiae and metazoanOrc4. Unlike S. cerevisiae ORC, which utilizes all AAAþdomain-containing Orc subunits for DNA binding (Lee &Bell, 1997), S. pombe Orc4 is necessary and sufficient forinitial origin engagement (Chuang & Kelly, 1999;Gaczynska et al., 2004; Kong & DePamphilis, 2001; Moonet al., 1999), and the AT-hook motifs of Orc4 are add-itionally required for viability (Chuang et al., 2002).Interestingly, S. pombe ORC shows a biphasic mechan-ism of DNA binding, with an initial, salt-sensitive DNAbinding event that precedes the formation of a salt-sta-ble form, a state that in turn can be further stabilizedby the addition of Cdc6 (Houchens et al., 2008). Thesefindings suggest that S. pombe ORC is recruited to chro-mosomes by the Orc4 AT-hook motif, which then leadsto stable DNA association through a mechanism thatlikely involves the encirclement of duplex DNA.

Analogous to S. pombe Orc4, metazoan Orc6 hasbeen shown to have a distinct DNA binding activity.Despite the absence of an ATPase or WH domain, exclu-sion of Orc6 from D. melanogaster ORC results in theloss of ATP-dependent DNA binding, the same effectobserved 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 (Balasovet al., 2007; Chesnokov et al., 2003), and mutation of theOrc6 TFIIB domain abolishes DNA binding (Liu et al.,2011). Although the Orc6 TFIIB domain was initially con-sidered unique to metazoans, subsequent sequenceanalysis indicates that the domain is conserved in fun-gal Orc6, but lacks specific DNA-binding elements(Bleichert et al., 2013), a finding consistent with S. cerevi-siae Orc6 being dispensable for origin recognition invitro (Chen et al., 2007; Lee & Bell, 1997). Given theavailable data, it seems likely that metazoan ORC uti-lizes Orc6 to loosely tether the complex to DNA in afunctionally analogous manner as S. pombe Orc4, andthat this action aids with the initial ORC recruitmentevent that precedes stable origin association. Thismodel begs the question, however, of how ORC isrecruited to chromosomes in species where Orc6 is alabile subunit (Dhar et al., 2001; Gillespie et al., 2001;Vashee et al., 2001). One possible answer is that chro-matin-bound Orc6 may function as a recruitment plat-form 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 forrecruiting and regulating ORC’s association with origins(Chakraborty et al., 2011). Early studies recognized thatORC played dual roles in replication and transcriptionalregulation (Bell et al., 1993; Leatherwood & Vas, 2003;Loo et al., 1995; Palacios DeBeer et al., 2003), and anumber of transcriptional regulators have been shownto bind ORC, including HP1, E2F, HMGA1a and Sir1(Bosco et al., 2001; Fox et al., 1997; Gardner et al., 1999;Pak et al., 1997; Prasanth et al., 2004, 2010; Royzmanet al., 1999; Thomae et al., 2008; Triolo & Sternglanz,1996). Although chromatin accessibility and transcrip-tional programs are clearly regulated in an ORC-dependent manner (Bell et al., 1993; Bose et al., 2004;Chesnokov, 2007; Foss et al., 1993; Fox et al., 1995;Huang et al., 1998; Loo et al., 1995; Micklem et al., 1993;Pak et al., 1997; Shor et al., 2009), whether chromatinbound transcriptional regulators also direct ORC originusage is less clear. Certain cases have been investigatedin some detail; for example, ORC can be targeted tochromatin-bound HMGA1a, where it directs assembly ofreplication complexes (Thomae et al., 2008), while gen-ome-wide analysis of metazoan origins reveals ORCassociation with HP1 sites (Cayrou et al., 2011). Thus,transcriptional regulators represent an additional meansof tethering ORC to chromosomes and likely affect ORCfunction in replication.

Eukaryotic ORC is also directly recruited to histonesby the N-terminal Bromo-Adjacent Homology (BAH)

114 M. W. PARKER ET AL.

Page 11: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

domain in Orc1 (Noguchi et al., 2006; Muller et al., 2010;Zhang et al., 2002). This element selectively bindsH4K20me2, an interaction that directs replication licens-ing (Kuo et al., 2012). Other histone modifications alsohave been found to direct origin usage (such asH4K20me1, H3K4me2/3, and H4 acetylation), but howthese effect ORC positioning or downstream licensingreactions remains unclear (Aggarwal & Calvi, 2004;Costas et al., 2011; Eaton et al., 2011; Knott et al., 2009;Liu et al., 2012b; Miotto & Struhl, 2010; Rice et al., 2002;Vogelauer et al., 2002; Tardat et al., 2010). A factorknown as ‘ORCA’ (for Origin Recognition ComplexAssociated) additionally has been reported to directlyrecruit human ORC to chromosomes, as well as tobind to other initiation factors that promotereplication licensing and S-phase progression(Shen et al., 2010, 2012). Conversely, the assembly ofthe replication machinery at telomeres has beenreported to be controlled by an interaction betweenORC and the TRF2 subunit of the shelterin complex(Atanasiu et al., 2006; Deng et al., 2009; Higa et al., 2016;Tatsumi et al., 2008).

Collectively, multiple lines of data demonstrate thatlocalizing ORC to DNA prior to productive originengagement (i.e. Cdc6-dependent DNA encircling) is acritical step in replication initiation, and that manydiverse recruitment mechanisms are sufficient todemarcate a replication start site. Consistent with thisidea is the demonstration that a fusion between ORCand the Gal4 DNA binding domain can result in the ini-tiation of DNA replication on a plasmid containing atandem array of Gal4 binding sites (Takeda et al., 2005).In light of the many mechanisms that can facilitate ORCrecruitment, an important future direction will be tounderstand how different recruitment pathwayscooperate or antagonize each other in specifying sitesof replisome assembly.

The minichromosome maintenance (MCM)complex

Following the active designation of an origin byORC, the complex next facilitates initiation at thesesites by loading the heterohexameric MCM2–7 heli-case onto DNA. Studies into MCM2–7 function haverevealed a surprisingly complex enzyme that appearscapable of harnessing ATP to promote both themelting of double-stranded DNA and translocation ofsingle-stranded DNA substrates. Although precisemechanisms have yet to be elaborated, a picture ofhow specific MCM2–7 elements couple ATP turnoverto DNA remodeling and movement is beginning toemerge.

Identification of the MCM2–7 complex

MCM proteins were originally identified in geneticscreens that aimed to uncover factors required for repli-cation and cell cycle progression in yeast (Chen et al.,1992; Hennessy et al., 1990; Maine et al., 1984; Moiret al., 1982; Yan et al., 1991). Homologs were subse-quently 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 shownto have an essential role in DNA replication.Investigation into MCM function converged with studiesby Laskey and Blow, who identified a replication licens-ing factor (RLF) that restricted replication to once percell cycle by nuclear exclusion until after nuclear enve-lope breakdown (Blow & Laskey, 1988; Blow, 1993).Indeed, MCM proteins, such as the RLF, showed redistri-bution from the cytosol to nucleus upon completion ofmitosis (Hennessy et al., 1990), and immunodepletion ofMCM proteins inhibited replication (Kubota et al., 1995;Madine et al., 1995). Importantly, a purified RLF fractionwas found to resolve into two separate factors, RLF-Mand RLF-B, with RLF-M containing multiple proteins thatcross-reacted with MCM antibodies (Chong et al., 1995).

Purification of RLF-M, as well as co-immunoprecipita-tion studies, revealed that MCM proteins reside withinlarge multimeric assemblies with other MCM2–7 familymembers (Burkhart et al., 1995; Chong et al., 1995, 1996;Kubota et al., 1995; Kimura et al., 1995; Madine et al.,1995; Musahl et al., 1995; Romanowski et al., 1996a).The predominant assembly in vivo is an MCM2–7 heter-ohexamer, although low levels of subassemblies havebeen reported to exist that may represent intermediates(Adachi et al., 1997; Ishimi, 1997; Lee & Hurwitz, 2000;Prokhorova & Blow, 2000; Su et al., 1996; Thommeset al., 1997). As with Orc1–5, MCM proteins are predi-cated upon a conserved AAAþ ATPase element (Iyeret al., 2004; Koonin, 1993). However, unlike Orc1–5,which show differing levels of active site conservation,the six subunits in the MCM2–7 complex each containthe complete set of catalytic residues expected to benecessary for supporting ATP hydrolysis.

MCM architecture

The basic architecture of an MCM protein is conservedacross archaea and eukaryotes. MCMs contain a centralAAAþ ATPase fold flanked by conserved N- and C-ter-minal domains (termed “NTD” and “CTD”, respectively).Structural analyses of MCMs have revealed three con-served subdomains within the NTD, allowing subdiv-ision of this element into the NTD-A, -B and -C regions(Figure 4(A)) (Bae et al., 2009; Brewster et al., 2008;

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 115

Page 12: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Fletcher et al., 2003; Fu et al., 2014; Li et al., 2015; Liuet al., 2008). NTD-C, the most highly conserved NTDsubdomain, possesses an oligonucleotide/oligosacchar-ide-binding (OB) fold. Within an MCM hexamer, interac-tions between adjacent NTD-C elements serve as theprimary interface for subunit assembly, with this inter-action alone sufficient to facilitate hexamerization ofthe NTD (Fletcher et al., 2003; Kasiviswanathan et al.,2004). Although NTD-A is the least conservedN-terminal subdomain, it forms a helical bundle withsimilarity to helix-turn-helix type DNA binding proteins(Costa et al., 2008). The NTD-B element comprises aZn-finger motif that, with the exception of MCM3, isuniversally conserved in MCM2–7 (Li et al., 2015), andthat facilitates higher order organization of single NTDhexamers into a conserved, head-to-head double hex-amer (Figure 4(B)) (Chong et al., 2000; Costa et al., 2014;Evrin et al., 2009; Fletcher et al., 2005; Kelman et al.,1999; Li et al., 2015; Remus et al., 2009; Shechter et al.,2000).

The AAAþ domain is the most highly conservedregion across MCM homologs. The MCM AAAþ fold

contains three unique insertions relative to prototypicalAAAþ proteins, and falls within the pre-sensor II (PSII)insert clade of the ATPase superfamily (Erzberger &Berger, 2006; Iyer et al., 2004). One such insertion is dis-tinctive in that it remodels a portion of the AAAþ cas-sette to orient a catalytic amino acid known as thesensor II motif in trans with the active site of an adja-cent protomer (Bae et al., 2009; Moreau et al., 2007),rather than in cis, which serves as the more usualarrangement in AAAþ oligomers. By comparison, theother two insertions are found in many other AAAþsubgroups and consist of two b–hairpins termed thepre-sensor I (PSI)-insert and the helix 2 (H2)-insert.Within the context of the hexamer, these b-hairpins arepositioned within the central channel and are integralto helicase mechanism (Figure 5). With regard to theMCM catalytic centers, a recent high-resolution cryo-electron microscopy structure of a full-length S. cerevi-siae MCM2–7 double hexamer reveals significant struc-tural variability between the six radially arrangedATPase sites, with a catalytically competent conform-ation but unequal nucleotide occupancy observed for

Figure 4. MCM architecture. (A) MCM homologs are characterized by three domains: NTD, AAAþ, and CTD. The NTD can be sub-divided into NTD-A (a small helical bundle), NTD-B (Zn-finger), and NTD-C (OB fold). The CTD forms a WH domain (for AAAþ andNTD, PDB¼ 3F9V; for WH domain, PDB¼ 2KLQ). (B) Two physiologically relevant MCM oligomers have been observed, a hexamerthat is formed by lateral interactions between the AAAþ and NTD domains of adjacent protomers, and a double hexamer thatis 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 viewthrough the central cavity and the radial arrangement of eukaryotic MCM2–7 subunits (the double hexamer was built from twocopies of PDB¼ 3JA8 fit to the EM density map EMD-6338 (Li et al., 2015)). A color version of this figure is available at www.tandfonline.com/ibmg.

116 M. W. PARKER ET AL.

Page 13: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

four of the sites (Li et al., 2015). This variability is con-sistent with the unequal ATPase activity observed forthe six MCM2–7 active sites (Bochman et al., 2008) andwith the nonequivalent function of these sites acrossthe different stages of helicase recruitment, loading andactivation (Coster et al., 2014; Ilves et al., 2010; Kanget al., 2014).

Insofar as the MCM CTD, solution structures haveshown this region to adopt a canonical winged-helix(WH) domain (Figure 4(A)) (Liu et al., 2012a; Wei et al.,2010; Wiedemann et al., 2015), which based on theweak or absent CTD density in structures of many full-length MCM proteins, appears flexibly tethered to theAAAþ core. Although the position of this domain withrespect to the hexamer was at first unclear, recent struc-tural 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-strandedDNA weakly (Pucci et al., 2007); however, what role thisactivity plays in vivo, and whether the WH domains ofthe eukaryotic MCMs can also bind DNA, are open ques-tions. The majority of data reported thus far suggeststhat, at least in archaea, the MCM WH domain functionsto allosterically modulate the ATP hydrolysis rate of theAAAþ domain (Barry et al., 2007; Jenkinson & Chong,2006; Wiedemann et al., 2015). In addition, this regioncan serve as a protein/protein interaction site, with theWH domains of eukaryotic MCM3 and MCM6 havingbeen shown to engage Cdc6 and Cdt1, respectively(Frigola et al., 2013; Liu et al., 2012a; You & Masai, 2008).Interestingly, recent structural analyses of the activated

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 AAAþ domain(PDB ¼3JA8, chain 2). (B) In the context of a hexamer, the MCM functional elements (excepting the external b-hairpin) line a cen-tral cavity through which DNA translocates (modeled after PDB ¼3JA8, chains 4, 6, and 7). (C) Symbol key. (D) Detailed functionaldescription for each MCM element known to contribute to activity. A color version of this figure is available at www.tandfonline.com/ibmg.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 117

Page 14: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

yeast helicase (the CMG complex) have revealed thatthe Mcm5 and Mcm6WH domains partially occlude theMcm2–7 central channel, thus positioning them to per-haps function in the translocation process (Yuan et al.,2016). Despite this observation, the precise role of theMCM WH domains remains to be determined.

Although the ATPase activity of MCM proteins local-izes exclusively to the core AAAþ domain, hexamer for-mation is facilitated by both the AAAþ and NTDelements in the order MCM5!MCM3!MCM7!MCM4!MCM6!MCM2 (Costa et al., 2011; Crevel et al.,2001; Davey et al., 2003; Schwacha & Bell, 2001 ). A loopwithin each NTD-A OB-fold, termed the allosteric com-munication loop (ACL), contacts the AAAþ domain ofits adjacent protomer, further stabilizing the hexamer(Li et al., 2015; Miller et al., 2014). Collectively, theseinteractions facilitate the formation of a particle consist-ing of two stacked rings (a AAAþ tier and an NTD tier).In single and double MCM2–7 hexamers, the centralchannel is sufficiently large enough to accommodatedouble-stranded DNA (Figure 4(B)) (Costa et al., 2011,2014; Evrin et al., 2009; Li et al., 2015; Lyubimov et al.,2012; Remus et al., 2009); later, in an unknown series ofevents, the helicase undergoes an isomerization reac-tion that appears coupled to the extrusion of one of thetwo strands through a natural discontinuity betweenthe MCM2/5 subunits (Bochman & Schwacha, 2008;Bruck & Kaplan, 2015b; Costa et al., 2011), permittingtranslocation along single-stranded DNA (Fu et al., 2011;Ilves et al., 2010; Ishimi, 1997; Kelman et al., 1999;Shechter et al., 2000).

MCM mechanism in origin melting and processiveunwinding

The MCM2–7 complex matures through a variety ofintermediates before being incorporated into the activehelicase present at replication forks. Within theMCM2–7 lifecycle, two stable complexes are observed:an origin-bound double hexamer (Evrin et al., 2009;Remus et al., 2009) and a replication fork-associated sin-gle hexamer (Duzdevich et al., 2015; Fu et al., 2011).Interestingly, the double hexamer serves not only as aplatform for recruiting helicase-activating factors butalso is likely responsible for the initial origin meltingevent (Bochman & Schwacha, 2015; Li et al., 2015; Sunet al., 2014 ). Thus, the MCM complex must utilize theATP-hydrolysis-driven repositioning of its DNA-bindingelements to carry out two very distinct tasks, originopening as a double hexamer and processive DNAunwinding as a single hexamer. Although the MCM2–7DNA-binding elements are relatively well-defined, it iscurrently unclear whether the two functionalities of the

helicase require an overlapping or mutually exclusiveset of protein interactions with DNA.

The MCM2–7 hexamer has multiple tiers of DNA-binding elements that are positioned to engage DNApassing through the central pore (Figure 5). The NTD-BZn-fingers are situated at one end of the hexamer,forming a skirt of DNA-binding elements that run paral-lel to and extend the central channel. This domain isrequired for double hexamer formation, and mutationor deletion of the archaeal NTD-B reduces DNA bindingand results in loss of helicase activity (Kasiviswanathanet al., 2004; Poplawski et al., 2001; Pucci et al., 2004).The Zn-finger motifs of eukaryotic MCM2–7 also facili-tate double-hexamer formation (Li et al., 2015) and rep-resent a putative DNA-binding region, but the role ofthis domain in helicase activity has not been investi-gated directly. The Zn-finger motif of S. cerevisiae Mcm2and Mcm5 are required for cellular proliferation, sug-gesting an essential and conserved function for thisregion (Dalton & Hopwood, 1997; Yan et al., 1991).

As one moves through the central MCM2–7 pore,from the NTD to the CTD, the next DNA-binding ele-ments encountered are the b–hairpin insertions withinthe OB-fold (NTD-C). The NTD-C b–hairpin constricts thecentral channel (Fletcher et al., 2003; Li et al., 2015) and,in archaea, is electropositive in nature. Mutation of basicresidues within this loop and on the adjacent, pore-lin-ing surface of the OB-fold dramatically reduce bothDNA binding and helicase activity in archaeal MCM(Fletcher et al., 2003, 2008; McGeoch et al., 2005; Pucciet al., 2004). The NTD-C b-hairpin is highly variable inlength and composition within the eukaryotic MCM2–7family, and while a majority are enriched in basic resi-dues, this conservation is not universal. Nevertheless,the NTD-C region appears critical for eukaryotic MCMfunction, as S. cerevisiae strains harboring mutationswithin this b–hairpin of Mcm5 show an increased rateof minichromosome loss (Leon et al., 2008). This effectcan be repressed by the addition of multiple ARSsequences, which suggests that origin recruitment, butnot helicase activity, is defective in this particularmutant.

In addition to the NTD-C b–hairpin, the crystal struc-ture of an archaeal MCM NTD bound to single-strandedDNA has revealed a second DNA-binding region withinthe NTD-C, the MCM-single-stranded-DNA binding motif(MSSB) (Froelich et al., 2014). The MSSB is formed pri-marily by two positively charged residues that extendfrom the OB-fold and, interestingly, bind single-stranded DNA in an orientation perpendicular to thelong axis of the central channel. The MSSB DNA-bindingresidues are conserved among three consecutive subu-nits of the eukaryotic MCM2–7 complex (MCM4, MCM6

118 M. W. PARKER ET AL.

Page 15: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

and MCM7), but not the other subunits. It has been sug-gested that the MSSB plays a role in DNA melting, work-ing against a DNA pumping action of the ATPasedomains to induce topological strain that encouragesstrand separation (Froelich et al., 2014). Thus, the MSSBmay have a role in the early steps of initiation, consist-ent with the defects observed for S. cerevisiae MSSBmutants in in vitro loading reactions (Froelich et al.,2014); whether or how this motif contributes to helicasefunction during elongation is unknown.

The AAAþ PSI b–hairpin insertion constitutes themost C-terminal DNA-binding element. This pore-liningloop shows the highest level of conservation of the vari-ous DNA-binding elements discussed thus far, and con-tains an invariant lysine that projects toward the centralpore. The PSI b-hairpin has been compared to the trans-location b–hairpin of SF3 helicases (e.g. the papillomavi-rus E1 protein and the SV40 Large T-antigen), which inthe context of a hexamer forms a vertically aligned,right-handed staircase that tracks the DNA backbonewith a conserved lysine (Enemark & Joshua-Tor, 2006,2008). Consistent with a critical role in helicase function,mutation of the PSI lysine in archaeal MCM weakensDNA binding and ATP hydrolysis, and fully abrogateshelicase activity (McGeoch et al., 2005); alteration of thisresidue within the context of the D. melanogaster CMGsimilarly abolishes helicase function (Petojevic et al.,2015). Interestingly, genetic analysis of the S. cerevisiaePSI b–hairpin lysines has highlighted nonequivalentfunctions for this residue in the six Mcm2–7 homologs,with only the mutation of this residue in Mcm3 abolish-ing an ability to complement deletion strains (Lamet al., 2013; Ramey & Sclafani, 2014).

How are the activities of the MCM2–7 DNA-bindingelements functionally coordinated with ATP turnoverto achieve the different functionalities observed forthe helicase? At least two other loops seem to facili-tate communication between the AAAþ domain andthe DNA-binding elements lining the central channel.One is a motif known as the H2-insert, which sits inthe MCM AAAþ fold and forms an extended loopthat junctions the ATPase and NTD tiers, creating apore-lining feature at the interface (Bae et al., 2009;Brewster et al., 2008; Li et al., 2015). Notably, analysisof the current data suggests that ATP-dependentrepositioning of the H2-insert may switch the helicasebetween an origin melting conformation to one thatfacilitates processive unwinding. Despite a low levelof sequence conservation, the length of the H2-insertis fully conserved in archaea and all MCM2–7 homo-logs and is sufficiently long to directly contact theMSSB. Interestingly, the H2-insert is rich in chargedamino acids and alignment of the six yeast Mcm

proteins reveals two fully conserved acidic residuesthat, in the context of the double hexamer, are posi-tioned to shield the MSSB from binding DNA (Figure5). Consistent with this proposal, the H2-insert dra-matically affects DNA binding in the archaeal MCMcomplex, with removal of the loop enhancing MCMaffinity for both double and single-stranded DNA(Jenkinson & Chong, 2006). The position of the H2-insert is thought to be modulated in an ATP-depend-ent fashion, such that the presence of ATP results ina more buried state for this loop (Jenkinson & Chong,2006). In conclusion, the available data suggest a crit-ical role for the H2-insert in transitioning the helicasebetween different functional states.

In addition to H2-insert-dependent crosstalkbetween MCM subunits, the ACL of NTD-C projectsupward from the OB-fold and toward the AAAþdomain of an adjacent protomer, where it is sand-wiched between the H2-insert and PSI b–hairpin(Barry et al., 2009; Li et al., 2015; Sakakibara et al.,2008). Like the other functional motifs within theNTD, the ACL is absolutely required for helicase activ-ity. However, the ACL does not affect DNA-bindingactivity, but instead modulates the ATPase activity ofthe AAAþ domain; mutation of ACL residues or dele-tion of the region results in markedly reduced ATPaseactivity and inhibition of duplex unwinding (Barryet al., 2009; Sakakibara et al., 2008). Like the H2-insert,the ACL position with respect to the AAAþ domain isaltered under different nucleotide-bound states(Barry et al., 2009).

Overall, the DNA-binding elements within the MCMNTD appear to play a critical function in both originmelting and unwinding. Although functional datapoint to an ATP-controlled connection between thepositional status of specific DNA binding elementswithin the MCM2–7 ring and their ability to grasp orrelease substrate, direct observation of these elementsin either a DNA melting or translocation mode is cur-rently lacking. Recent electron microscopy models ofthe activated eukaryotic MCM2–7 helicase in the con-text of the CMG have revealed that there exists con-formational coupling between the NTD and CTD tier,suggestive of coordinated action between theseregions during processive unwinding (Abid Ali et al.,2016; Yuan et al., 2016). NTD-A also exhibits conform-ational dynamics and has been seen to rotate mark-edly away from the hexamer axis, a movement thatmay be coordinated through direct binding of DNAto the NTD-A (Chen et al., 2005; Costa et al., 2008;Fletcher et al., 2003; Hoang et al., 2007; Liu et al.,2008). As additional substrate-bound structures aresolved for the archaeal and eukaryotic MCM

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 119

Page 16: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

homologs, exciting insights into its DNA remodelingand motor mechanisms are certain to emerge.

Loading of the replicative helicase

The defining step in eukaryotic replication licensing isloading of MCM2–7 onto duplex DNA into a stablehead-to-head double hexamer (Figure 1) (Evrin et al.,2009; Gambus et al., 2011; Remus et al., 2009). To attainthis state, single MCM2–7 hexamers are sequentiallyloaded by an interaction with ORC·Cdc6 that is cha-peroned by the Cdc10-dependent transcript 1 (Cdt1)protein. Structural characterization of multiple inter-mediates, together with the ability to reconstitute andstudy loading in vitro, are revealing how the stepwiseand carefully orchestrated exchange of pre-RC factors atorigins underlies this complex reaction (Figure 6).

Mechanism of helicase recruitment to origins

Maturation of the pre-RC occurs in a sequential fashionduring the G1 phase of the cell cycle (Evrin et al., 2009;Gillespie et al., 2001; Romanowski et al., 1996b; Remuset al., 2009; Ticau et al., 2015; Tsakraklides & Bell, 2010;Tsuyama et al., 2005). Demarcation of replication startsites by ORC represents the inaugurating event, and

permits the recruitment of and stable association withCdc6 (Cocker et al., 1996; Coleman et al., 1996; Donovanet al., 1997; Liang et al., 1995; Seki & Diffley, 2000;Tsuyama et al., 2005). In an in vitro setting, Cdc6 regu-lates the fidelity of origin selection by two distinctmechanisms. First, Cdc6 helps to restrict the initiatorfrom acting at illegitimate origins by sequestering thefree initiator (thereby lowering the effective concentra-tion of the initiator to increase origin specificity)(Duzdevich et al., 2015), and by triggering the dissoci-ation of ORC from nonorigin DNA (Mizushima et al.,2000). Second, Cdc6 ATPase activity is enhanced whenbound to non-ARS sequences, which triggers the dis-sociation of Cdc6 from the initiator (Speck et al., 2005;Speck & Stillman, 2007). Ultimately, the secure associ-ation between Cdc6 and ORC at origins primes the initi-ator for helicase recruitment and loading (Donovanet al., 1997; Feng et al., 2000; Perkins & Diffley, 1998;Rowles et al., 1996; Tsuyama et al., 2005); this samecomplex will later facilitate the release of the doublehexamer for subsequent activation at the onset of Sphase (Chang et al., 2015).

The MCM2–7 helicase must first be recruited to ori-gins to initiate the loading reaction. Notably, the heli-case is sensitive to solution conditions and showssubstantial conformational heterogeneity (Bochman &

Figure 6. MCM2–7 complex loading and maturation into the CMG. Two sequential rounds of helicase recruitment and loading atorigins are required for building an MCM2–7 double-hexamer. For both hexamers, DNA is threaded into the central channelthrough 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 hex-amer loads, both Cdc6 and Cdt1 are released. Rebinding of Cdc6 to ORC primes the system for recruiting and loading a secondhexamer in the opposite direction as the first, an event that has been proposed to be controlled by ORC·Cdc6, but templated bythe first MCM2–7 hexamer. Cdt1 and Cdc6 recruitment and ejection are required for both loading events. Phosphorylation of thedouble hexamer and other initiation factors by CDK and DDK facilitate origin melting, GINS and Cdc45 recruitment/assembly, DNAstrand extrusion, and activation of the helicase for DNA unwinding. A color version of this figure is available at www.tandfonline.com/ibmg.

120 M. W. PARKER ET AL.

Page 17: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Schwacha, 2008; Costa et al., 2011; Gomez-Llorenteet al., 2005), a feature that is pertinent for understand-ing the mechanism of association with the origin-boundinitiator. Biochemical and structural analyses of theMCM2–7 complex has revealed that, in the absence ofATP, archaeal, fungal and metazoan complexes canadopt a cracked ring architecture that bears a discon-tinuity, or “gate”, between MCM2 and MCM5 (Bochman& Schwacha, 2008; Boskovic et al., 2016; Costa et al.,2011; Lyubimov et al., 2012; Samson et al., 2016).Nucleotide binding constricts the ring and favors theformation of MCM2/5 interactions across the ring break(Samel et al., 2014), but does not appear to irreversiblyclose the gate in single MCM2–7 hexamers (Costa et al.,2011; Lyubimov et al., 2012). Thus, MCM2–7 may notrequire active ring-opening prior to loading, but insteadmay 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 andmetazoa (Claycomb et al., 2002; Devault et al., 2002;Maiorano et al., 2000; Nishitani et al., 2000; Whittakeret al., 2000). However, the role of this factor has beenenigmatic. Although Cdt1 was initially thought to berequired for helicase recruitment to ORC (Asano et al.,2007; Chen et al., 2007; Chen & Bell, 2011), recent datasuggest that MCM2–7 is recruited by direct interactionswith the initiator (Fernandez-Cid et al., 2013; Frigolaet al., 2013). With the exception of Orc6, all initiationfactors discussed thus far possess a central AAAþATPase unit; Cdt1 represents an additional non-AAAþprotein involved in the process. Cdt1 was first identifiedin S. pombe, and through an ability to induce re-replica-tion in the absence of cell division, was recognized asan essential and potent initiator of replication(Hofmann & Beach, 1994; Nishitani et al., 2000). Cdt1 isuniversally conserved in metazoa but is inconsistentlyfound in other eukaryotic supergroups (Aves et al.,2012; Maiorano et al., 2000; Tada et al., 1999; Whittakeret al., 2000). The weak conservation of Cdt1 acrossspecies aggravated initial attempts to identify an S.cerevisiae homolog, but eventually the TAH11(Topoisomerase-A hypersensitive 11) protein was identi-fied as the budding yeast counterpart (Devault et al.,2002; Tanaka & Diffley, 2002). TAH11 was first isolatedin genetic screens for genes that showed syntheticlethality with a mutant topoisomerase I allele (Fiorani &Bjornsti, 2000), and proteomic studies in human cellshave suggested there exists an interaction betweenCdt1 and both topoisomerase I and topoisomerase IIa(Sugimoto et al., 2008). Archaea also contain a Cdt1-related protein, termed WhiP; however, this factorappears to uniquely function in the Orc-independent

assembly of pre-RCs (Robinson & Bell, 2007; Samsonet al., 2013).

Interestingly, Cdt1 has markedly diverged between S.cerevisiae and metazoa, and functional analyses havesuggested that there exist differing mechanisms forCdt1-dependent action across species. All Cdt1 homo-logs are built from a conserved C-terminal pair of WHdomains (Khayrutdinov et al., 2009; Lee et al., 2004). Bycontrast, 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 pI¼ 10.5) thatis predicted to be unstructured, the S. cerevisiae N-ter-minus has an acidic pI (pI¼ 5.1) and possesses sufficientsecondary structure to likely constitute an ordereddomain (modeling using Phyre2 (Kelley et al., 2015) sug-gests that this region corresponds to a catalyticallydefunct oxygenase fold). Interestingly, an S. cerevisiaeCdt1 construct lacking the nonconserved N-terminaldomain fails to load the helicase into an activation-com-petent conformation (Fernandez-Cid et al., 2013; Takara& Bell, 2011), whereas the N-terminus of metazoa Cdt1is dispensable for the loading reaction (Ferenbach et al.,2005).

The factors with which Cdt1 stably associates alsoevince species-specific differences. In S. cerevisiae, Cdt1and Mcm2–7 form a tight complex and exhibit inter-dependent nuclear import (Kawasaki et al., 2006; Remuset al., 2009; Takara & Bell, 2011; Tanaka & Diffley, 2002);Cdt1 similarly helps maintain the structural integrity ofthe Mcm2–7 complex (Frigola et al., 2013; Wu et al.,2012). Conversely, Xenopus Cdt1 and MCM2–7 are bio-chemically separable initiation factors, with immunode-pletion of MCM3 from Xenopus egg extracts removingMCM2–7 but having no effect on Cdt1 (Maiorano et al.,2004) (it is worth noting that mouse Cdt1 has beenreported to form a complex with MCM2–7 in vitro (You& Masai, 2008), but the stability of this assembly has notbeen investigated). Multiple explanations could accountfor the differences in behavior between S. cerevisiae andXenopus Cdt1; for example, differences in the relativeaffinities of Cdt1 for the MCM2–7 complex could reflectan alteration to the sequential assembly of the pre-RCfactors in metazoa, such that Cdt1 associates first withORC·Cdc6 rather than with the helicase (Maiorano et al.,2000; Waga & Zembutsu, 2006). Finally, metazoa con-tain an additional Cdt1-binding protein, Geminin, whichnegatively regulates licensing by inhibiting the Cdt1-dependent loading of the helicase into a stable doublehexamer (Edwards et al., 2002; Wohlschlegel et al., 2000;Wu et al., 2014a; Yanagi et al., 2002).

Despite the differences in Cdt1 structure and its pre-ferred binding partners between budding yeast andmetazoans, multiple studies have identified MCM6

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 121

Page 18: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

as the primary Cdt1 interaction site on the helicasehexamer (Ferenbach et al., 2005; Liu et al., 2012a; Weiet al., 2010; Wu et al., 2012; Yanagi et al., 2002; Zhanget al., 2010) (mouse Cdt1 also interacts with MCM2 (You& Masai, 2008)). Cdt1/MCM6 contacts are facilitated bythe C-terminal WH domain of each protein (Ferenbachet al., 2005; Jee et al., 2010; Khayrutdinov et al., 2009;Takara & Bell, 2011; Teer & Dutta, 2008; Wei et al., 2010),with structural studies of the human proteins revealingthat a helical extension which projects from theCdt1WH region engages the helix-turn-helix motif ofthe MCM6 C-terminus (Liu et al., 2012a; You & Masai,2008). Mutagenesis of residues in the interaction surfaceabolishes helicase loading and DNA replication in bud-ding yeast, suggesting that the MCM6-Cdt1 mechanismof binding is widely conserved (Liu et al., 2012a).Despite this contact, in vitro binding studies have sug-gested that there exist additional undefined sites ofinteraction between Cdt1 and MCM2–7 (Fernandez-Cidet al., 2013; Khayrutdinov et al., 2009).

Saccharomyces cerevisiae Cdt1 has also beenreported to bind Orc6 (Asano et al., 2007; Chen et al.,2007; Chen & Bell, 2011; Semple et al., 2006). This inter-action was initially proposed to facilitate helicaserecruitment by helping to tether MCM2–7 to theORC·Cdc6 complex (it is unclear whether metazoanCdt1 and Orc6 interact (Yanagi et al., 2002)); however,other work has indicated that S. cerevisiae Mcm2–7 con-tains elements that directly engage ORC·Cdc6 and thatare wholly sufficient for recruitment (Fernandez-Cidet al., 2013; Frigola et al., 2013). Cdt1’s participation inmodulating MCM/ORC·Cdc6 interactions is still underdebate, with one study noting that budding yeastCdt1 relieves an Mcm6-dependent autoinhibitorymechanism that prevents Cdt1-independent recruit-ment (i.e. deletion of an Mcm6 autoinhibitory domainresults in helicase recruitment in the absence of Cdt1)(Fernandez-Cid et al., 2013), and another showing thatCdt1 is fully dispensable for helicase recruitment (hereonly an interaction between Cdc6 and the extreme C-terminus of Mcm3 appears necessary for recruitment(Frigola et al., 2013). Consistent with the proposed roleof the eukaryotic MCM3 C-terminus, recent studies inarchaea have found that the MCM C-terminal WHdomain directly binds Orc to facilitate recruitment ofthe helicase to replication origins (Samson et al., 2016).

Together, the available data suggest that Cdt1 maybe dispensable for helicase recruitment, and that thisstep instead predominantly relies on interactionsbetween ORC·Cdc6-bound origins and MCM2–7. Whatthen might be the significance of Cdt1 and its inter-action with Orc6? Beyond a potential ability to relieveMCM6-dependent autoinhibition (Fernandez-Cid et al.,

2013), S. cerevisiae Cdt1 stabilizes Mcm2–7 at origins inan Orc6-dependent manner (Chen et al., 2007), suchthat in the absence of Cdt1, topologically linkedMcm2–7 hexamers fail to load (Chen & Bell, 2011).Notably, an N-terminal deletion construct of buddingyeast Cdt1 permits loading of double hexamers, but ina state that is incapable of being subsequently acti-vated (Takara & Bell, 2011). Thus, Cdt1 is required forevents downstream of recruitment, which appear toinclude the formation of an MCM2–7 double hexamerintermediate that is competent for replication. Overall,the molecular details underlying the role of Cdt1, and inparticular its Orc6 association, are in need of furtherstudy.

Loading the double hexamer

Ultimately, the initiator-dependent recruitment ofMCM2–7 to origins leads to the formation of a stablechromatin-bound MCM2–7 complex that, unlikeORC·Cdc6, is resistant to high-salt extraction (Bowerset al., 2004; Donovan et al., 1997; Edwards et al., 2002).Notably, the MCM2–7 loading reaction has been fullyreconstituted in vitro with purified proteins from S. cere-visiae (Evrin et al., 2009; Remus et al., 2009), which hasin turn allowed the use of single molecule and struc-tural techniques to understand both the nature of thesalt-stable complex and the events leading to its forma-tion. The details of the loading reaction discussed in thefollowing are derived exclusively from the studies on S.cerevisiae initiation factors; given some of the species-specific differences observed to date, it will be import-ant to reconstitute the loading reaction in vitro withother organisms for comparison.

Following the recruitment of Cdt1·Mcm2–7 toORC·Cdc6-bound origins, Mcm2–7 is deposited ontoDNA while both Cdc6 and Cdt1 are ejected from thepre-RC in an ATP hydrolysis-dependent fashion. Cdc6release occurs prior to Cdt1 both for single- and dou-ble-hexamer loading (Evrin et al., 2009; Kang et al.,2014; Remus et al., 2009; Ticau et al., 2015; Tsakraklides& Bell, 2010) (Figure 6). Singly-loaded Mcm2–7 hexam-ers result in the transient formation of a meta-stableORC·Cdc6·Cdt1·Mcm2–7 (OCCM) intermediate, whichcan be stabilized by the presence of a nonhydrolyzableATP analog. This property has been exploited to permitimaging of the complex by 3D cryo-electron microscopy(Sun et al., 2013). Within the OCCM, Cdt1 interfaces withthe N-terminal tier of Mcm2, Mcm6 and Mcm5. TheMcm3 C-terminus resides proximal to Cdc6, consistentwith the observed interaction between these regions invitro and the proposed role of this interaction in heli-case recruitment (Frigola et al., 2013; Sun et al., 2013).

122 M. W. PARKER ET AL.

Page 19: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Cdt1 and Orc6 make no visible contacts in the OCCMstructure; the N-terminal TFIIB domain of Orc6 also doesnot appear ordered. Notably, the Cdt1·Mcm2–7 hep-tamer is oriented such that the MCM C-terminaldomains (the region containing the AAAþ and WHdomains) abut the ORC WH domains (Bleichert et al.,2015), leaving the MCM NTD accessible. Although ana-lysis of the loading reaction in bulk has indicated thatsingly loaded Mcm2–7 hexamers are salt-sensitive (Evrinet al., 2009; Kang et al., 2014; Remus et al., 2009), singlemolecule analysis of the reaction reveals that nearly50% of the singly loaded helicases are salt-stable, sug-gesting that at this stage Mcm2–7 ring closure hasoccurred (Ticau et al., 2015). This discrepancy betweenbulk and single-molecule studies currently lacks explan-ation, although may arise from the relatively high pro-tein concentrations used in the bulk assays and/or bytheir limited kinetic sensitivity.

After the first loading event, Cdc6 re-associates withthe ORC·Mcm2–7 complex, forming a short-livedORC·Cdc6·Mcm2–7 (OCM) intermediate that is compe-tent for recruitment of a second Cdt1·Mcm2–7 hep-tamer (Fernandez-Cid et al., 2013; Ticau et al., 2015).At present, the mechanics by which a second hexameris recruited are unclear. It has been proposed that load-ing of the second hexamer requires equivalent interac-tions as the first, with the Mcm3 C-terminus beingrequired for both loading steps (Frigola et al., 2013).However, if loading the second helicase into a doublehexamer requires an interaction with ORC·Cdc6,this mechanism must involve some fairly complicatedacrobatics, mainly because the first loaded Mcm2–7complex should still be present to block the MCM-bind-ing surface on ORC, and also because the secondMcm2–7 complex must be placed at a location morethan �100Å from ORC’s initial binding site on DNA, inan inverted orientation from the first helicase. This spa-tial problem could be solved if two ORCs were used forthe loading of two helicases with opposing polarity;however, single-molecule analysis has demonstratedthat a single, stably bound ORC is sufficient to load adouble hexamer (Ticau et al., 2015). Because of thesesteric complications, an alternative mechanism hasbeen proposed wherein the first, loaded hexamer tem-plates loading of the second through a distinct mechan-ism that does not require direct ORC·Cdc6 interactions(Ticau et al., 2015). Consistent with this idea, the kineticsof loading the second hexamer are slower than the first(Fernandez-Cid et al., 2013), as is the release of Cdc6and Cdt1 (Ticau et al., 2015), suggesting different reac-tion intermediates are accessed.

While the model described above does not directlyaccount for the demonstrated role of the budding yeast

Mcm3WH domain in the loading of both the first andsecond Mcm2–7 hexamers (Frigola et al., 2013), previouswork has revealed that deletion of the Mcm3 C-ter-minus results in a dramatic loss of Mcm2–7 ATPaseactivity (Sun et al., 2014). Since proper ATPase functionby the Mcm2–7 hexamer is critical at all stages of theloading reaction in S. cerevisiae, this loss of activity inthe Mcm3WH domain mutant could account for theobserved defect in loading the second hexamer.Accordingly, Mcm2–7 complexes with individual activesite mutations also show defective recruitment (particu-larly Mcm2 Walker A and Mcm6 arginine fingermutants), Cdt1 release, and dramatically reduced levelsof salt-stable double hexamers (Coster et al., 2014; Kanget al., 2014). Although a marginal level of double-hex-amer formation can be achieved with particular MCMATPase mutants, these loaded complexes are generallydeficient for downstream activation events and cannotdrive DNA replication in vitro (Kang et al., 2014).Collectively, the present data indicate that ATP bindingand hydrolysis by S. cerevisiae Mcm2–7 is required fromthe early steps of helicase recruitment to the conclud-ing activation events. These results are difficult torationalize with the observation that Xenopus MCM6and MCM7 Walker A mutants are loaded normally(Ying & Gautier, 2005), but this finding may again reflectspecies-specific differences.

Although double-hexamer loading is criticallydependent on the ATPase activity of Mcm2–7, ATPhydrolysis by Cdc6 and ORC perform additional rolesthat are consistent with a requirement for their ATPaseactivity in vivo (Bowers et al., 2004; Schepers & Diffley,2001; Takahashi et al., 2002; Weinreich et al., 1999;Zwerschke et al., 1994). ATP hydrolysis by Orc1 is dis-pensable for the loading reaction (Evrin et al., 2013;Fernandez-Cid et al., 2013), but resets the initiationmachinery to allow reiterative loading of MCM doublehexamers at origins (Bowers et al., 2004). By compari-son, Cdc6 has been reported to function as a qualitycontrol factor, whose ATPase activity ensures thatdefective or incomplete Mcm2–7 complexes arereleased from origins (Coster et al., 2014; Frigola et al.,2013; Kang et al., 2014). Another major function of Cdc6ATPase activity appears to be in the final step of pre-RCformation, where it is responsible for releasing success-fully loaded double hexamers from ORC, an event thatenables the downstream activation of the helicase(Chang et al., 2015). Loss of Cdc6 ATPase activity alsoresults in inefficient Cdt1 release (Randell et al., 2006),although ATP turnover by Mcm2–7 appears primarilyresponsible for Cdt1 ejection (Coster et al., 2014; Kanget al., 2014). Overall, Cdc6 functions as a linchpin factorcritical for multiple events in the loading process.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 123

Page 20: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Helicase activation

Although a double MCM2–7 hexamer is loaded duringpre-RC formation, at replication forks this structure isdissolved to yield two single MCM2–7 hexamers thatmove in opposite directions (Ticau et al., 2015; Yardimciet al., 2010). On their own, single MCM2–7 hexamerspossess little or no helicase activity (Ishimi, 1997; Ilveset al., 2010; Kaplan et al., 2003; Lee et al., 2000; Youet al., 2003). Initially, this observation was difficult torationalize with the strong evidence that pointedtoward a role for the complex as the replicative heli-case, but it is now appreciated that MCM2–7 is incorpo-rated into a larger Cdc45·MCM2–7·GINS (CMG) complexthat possesses robust helicase activity. Like pre-RCassembly, double hexamer dissolution and CMG forma-tion occur through a highly orchestrated and intercon-nected series of events that provide yet another meansfor regulating the DNA replication initiation reaction.

Identification of the Cdc45·MCM2–7·GINS (CMG)complex

The MCM2–7 genes are essential proteins in DNA repli-cation (Dalton & Whitbread, 1995; Gibson et al., 1990;Moir et al., 1982; Maine et al., 1984) and their phylogen-etic lineage to helicases led to early suggestions thatthese factors might function as nucleic acid motor pro-teins (Koonin, 1993). Together with work on archaealMCMs, which have been found to possess robust heli-case activity (Chong et al., 2000; Kelman et al., 1999;Shechter et al., 2000), these data indicated that theMCM2–7 complex likely served as the replicative heli-case in eukaryotes. Interestingly, initial in vitro studieswere unable to detect helicase activity from the eukary-otic MCM2–7 complex, although an MCM subcomplexcomposed of subunits (4,6,7)2 showed weak helicaseactivity (Ishimi, 1997; Kaplan et al., 2003; Lee et al., 2000;You et al., 2003). These data suggested that theMCM2–7 assembly might play a more limited role inreplication, such as origin melting. However, the persist-ence of MCM2–7 at replication forks (Aparicio et al.,1997), coupled with the observed S-phase arrest afterMCM2–7 inactivation (Labib et al., 2000; Pacek & Walter,2004; Shechter et al., 2004), provided irrefutable evi-dence for the role of the complex in elongation (Krudeet al., 1996; Todorov et al., 1995). Nevertheless, the dem-onstration of helicase activation remained elusive.

Eventually, conditions were identified that recoveredin vitro helicase activity for budding yeast Mcm2–7(Bochman & Schwacha, 2008). Roughly contemporan-eously, several lines of analysis began to reveal that thehelicase also existed within a larger macromolecular

assembly (Calzada et al., 2005; Gambus et al., 2006;Masuda et al., 2003; Moyer et al., 2006; Pacek et al.,2006). Three parallel studies – two using an unbiasedassessment of Cdc45 and GINS interacting factors and athird using a candidate-based approach to determinethe composition of the helicase stalled at replicationforks – have now demonstrated that MCM2–7 forms astable complex with Cdc45 and GINS (Gambus et al.,2006; Moyer et al., 2006; Pacek et al., 2006). Cdc45 andGINS are essential replication fork components (Dalton& Hopwood, 1997; Hopwood & Dalton, 1996; Kamimuraet al., 1998; Kanemaki et al., 2003; Masuda et al., 2003;Moir et al., 1982; Zou et al., 1997), which mimic thechromatin association pattern of MCM subunits (Kubotaet al., 2003; Takayama et al., 2003). Each MCM2–7 bindsa single copy of Cdc45 and one GINS tetramer to form astable, 11-protein complex termed the CMG (forCdc45·MCM2–7·GINS). The CMG has been shown to pos-sess robust helicase activity (Ilves et al., 2010; Moyeret al., 2006), and bioinformatics analyses have indicatedthat a full set of CMG factors are likely conserved fromarchaea to eukaryotes (Makarova et al., 2012), withrecent work demonstrating that, as observed for theeukaryotic accessory factors, archaeal Cdc45 and GINShomologs activate MCM in vitro and in vivo (Xu et al.,2016).

Regulating helicase activation

Appropriate recruitment of the factors necessary forCMG formation constitutes a highly regulated steptoward replisome formation. Consistent with the Cdc6-dependent release of the MCM2–7 double hexamerfrom ORC (Chang et al., 2015), helicase activation nolonger requires the function of either ORC or Cdc6 (Hua& Newport, 1998; Jares & Blow, 2000; Rowles et al.,1999; Walter, 2000; Yeeles et al., 2015). The newlyformed MCM2–7 double hexamer serves itself as a bind-ing platform and target for the Dbf4-dependent kinase(DDK) that, in collaboration with S-phase cyclin-depend-ent kinase (S-CDK), constitute the minimal set of kinasesneeded to regulate the recruitment of Cdc45 and GINS.Together, the activity of these two kinases coordinatethe assembly of a pre-IC that is poised for the activationand formation of a bidirectional replication fork(Figure 6) (Aladjem, 2007).

DDK is formed by the direct association of the Cdc7kinase with Dbf4, a protein co-factor that relieves Cdc7autoinhibition (Dowell et al., 1994; Jackson et al., 1993;Kitada et al., 1992) and that is regulated in a cell-cycle-dependent manner (Cheng et al., 1999; Oshiro et al.,1999; Nougarede et al., 2000). DDK phosphorylates mul-tiple MCM subunits (Cho et al., 2006; Lei et al., 1997;

124 M. W. PARKER ET AL.

Page 21: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Masai et al., 2006; Montagnoli et al., 2006; Sheu &Stillman, 2006; Weinreich & Stillman, 1999; Tsuji et al.,2006) and is required for initiating DNA replication(Chapman & Johnston, 1989; Hollingsworth & Sclafani,1990; Jiang et al., 1999a). Like DDK, S-CDK is requiredfor replication (Broek et al., 1991; Hayles et al., 1994)and exists in an inactive conformation by engagingSic1, a repressive protein that restricts S-CDK activityinto late G1/S phase (Donovan et al., 1994; Knapp et al.,1996; Mendenhall, 1993; Schneider et al., 1996; Schwobet al., 1994). Redundant regulation of S-CDK occurs bythe cell-cycle-dependent expression of activating cyclins(Abreu et al., 2013). It is currently unclear whether thereis a conserved sequential order for DDK and CDK activ-ity. In a X. laevis egg extract system, DDK activity is inde-pendent of CDK, and exposing the pre-RC first to CDKfully ablates DNA replication initiation (Jares & Blow,2000; Walter, 2000), suggesting that DDK acts first. Sucha consensus has not been found with S. cerevisiae pro-teins, where the sequential action of DDK and CDKdepends on the experimental setup (Heller et al., 2011;Nougarede et al., 2000) and has been found to beinconsequential in a fully in vitro reconstituted DNA rep-lication system (Yeeles et al., 2015).

DDK is recruited to Mcm2–7 through interactionswith Mcm4 and Mcm2 (Ramer et al., 2013; Sheu &Stillman, 2010) and targets specific MCM subunitswithin the context of a double, but not a single,MCM2–7 hexamer (Costa et al., 2014; Evrin et al., 2014;Francis et al., 2009; Kang et al., 2014; Sun et al., 2014).Notably, the MCM subunits responsible for recruitingDDK are nonadjacent within the context of a single hex-amer; however, the rotational offset within the doublehexamer results in the adjacent placement of MCM2and MCM4 from separate hexamers (Costa et al., 2014;Li et al., 2015; Sun et al., 2014), thus forming a compos-ite DDK-interaction interface. In addition, MCM2, MCM4and MCM6, which are all phosphorylated by DDK (Bruck& Kaplan, 2009; Cho et al., 2006; Lei et al., 1997; Masaiet al., 2006; Randell et al., 2010; Sheu & Stillman, 2006;Tsuji et al., 2006), reside in close proximity within thedouble hexamer and the interface between the twohexamers 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 itsMCM targets (Li et al., 2015; Sheu & Stillman, 2010; Sunet al., 2014). Collectively, the available data suggest thatthe DDK-dependent activation of MCM2–7 is temporallyregulated by selective recruitment of the kinase to andaction upon a double hexameric MCM2–7, a structurethat uniquely positions the relevant MCM subunits inclose 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 sensi-tize Mcm2–7 to DDK activity (Devault et al., 2008;Randell et al., 2010); however, DDK alone is required forthe activation of S. cerevisiae Mcm2–7 in vitro (Yeeleset al., 2015).

In S. cerevisiae, the DDK-dependent phosphorylationof Mcm2–7 is important for facilitating the Sld3 andSld7-chaperoned recruitment of Cdc45 onto the doublehexamer (Masai et al., 2006; Sheu & Stillman, 2006;Tanaka et al., 2011a; Yabuuchi et al., 2006). In humans,two proteins known as Treslin and MTBP (for MDM2binding protein) have been proposed to represent func-tional homologs of budding yeast Sld3 and Sld7,respectively (Boos et al., 2013; Bruck & Kaplan, 2015c;Kumagai et al., 2010; Matsuno et al., 2006; Sanchez-Pulido et al., 2010; Sansam et al., 2010). In vitro, Cdc45recruitment depends on the prior association of Sld3and Sld7 with the DDK-phosphorylated N-termini ofMcm2, Mcm4 and Mcm6 (Deegan et al., 2016; Fanget al., 2016); however, in vivo Sld3/7 and Cdc45 form astable complex and may be co-recruited to double hex-amers (Kamimura et al., 2001; Tanaka et al., 2011a,).Notably, in the absence of DDK, no additional initiationfactors are recruited to Mcm2–7 double hexamers.Conversely, in the absence of CDK activity, Cdc45 aswell as Sld3 and Sld7 are still efficiently recruited (Helleret al., 2011; Yeeles et al., 2015). Thus, recruitment ofCdc45 appears to rely exclusively on DDK.

How does DDK-dependent MCM2–7 phosphorylationfacilitate Cdc45 recruitment? Although we lack a com-plete mechanistic picture for DDK-dependent activation,three distinct MCM modifications have been identifiedthat result in DDK bypass, suggesting there exists acomplex role for DDK that likely involves a concertedchange within multiple MCM subunits. First, hyperphos-phorylation of the S. cerevisiae Mcm4 N-terminusrelieves an autoinhibitory function that this regionimposes on the helicase (indeed, deletion of the Mcm4N-terminus bypasses the DDK requirement) (Sheu &Stillman, 2006, 2010). Although a mechanism entailing asteric block to the Cdc45-binding site or the adoptionof a different nonpermissive orientation would satisfythis observation, within the atomic structure of the S.cerevisiae Mcm2–7 double hexamer the Mcm4 N-ter-minus is disordered, suggesting that autoinhibition isrelieved by an alternative mechanism (Li et al., 2015).Second, phosphorylation of Mcm4 and Mcm6 generatesa phosphopeptide-binding site for recruiting Sld3(phosphomimetic mutants of these MCM subunits canbypass the requirement for DDK) (Deegan et al., 2016).Finally, a mutation within budding yeast Mcm5, knownas the bob1 allele, has been found to bypass DDK-activated initiation (Hardy et al., 1997).

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 125

Page 22: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

In addition to DDK-dependent recruitment of Cdc45,Mcm2–7 activation requires an association with GINSand, for origin firing, with Replication Protein A (RPA),MCM10, Pola, Pole and CTF4 (van Deursen et al., 2012,Yeeles et al., 2015). Notably, CDK’s action on the grow-ing, chromatin-associated replication assembly, as wellas on replicative proteins yet to be integrated into thereplisome, results in the recruitment of the remaininginitiation factors that are needed for origin firing. In thisswarm of activity, CDK targets Cdc45, Sld2, Sld3 andSld7 (Heller et al., 2011; Masumoto et al., 2002;Muramatsu et al., 2010; Tanaka et al., 2007; Yeeles et al.,2015; Zegerman & Diffley, 2007), and, together with aprotein known as Dpb11, facilitates recruitment of amultiprotein assembly composed of GINS, Sld2 and Poleto the growing complex (Araki et al., 1995; Kamimuraet al., 1998, 2001; Muramatsu et al., 2010; Takayamaet al., 2003; Tanaka et al., 2013). Thus, DDK and CDKdrive formation of an active helicase by phosphorylat-ing key assembly factors that allow for the sequentialrecruitment of Cdc45 and GINS. For its part, MCM10 isrecruited by interactions with the CMG (Douglas &Diffley, 2016; Homesley et al., 2000; Wohlschlegel et al.,2002) and further appears to help stabilize the repli-some once formed (Gregan et al., 2003; Ricke &Bielinsky, 2004). Pole is likewise recruited directly to theCMG by an interaction with GINS (Araki et al., 1995;Muramatsu et al., 2010), whereas Pola is physicallycoupled to the helicase by CTF4 (Simon et al., 2014;Villa et al., 2016).

Many questions remain for the events that transitionthe growing MCM2–7 complex into a competent bi-dir-ectional replication fork. For example, it is currentlyunclear at what point and how origin melting occurs, orhow an MCM2–7 double hexamer, which encirclesduplex DNA, transitions into two uncoupled CMG par-ticles that encircle single-stranded DNA (Ticau et al.,2015; Yardimci et al., 2010). Interestingly, recent datasuggests that the process of origin melting and strandextrusion may be interdependent and temporally inter-twined with CMG assembly. Indeed, S. cerevisiae Sld2,Sld3 and Dpb11, which chaperone GINS into a complexwith the helicase, compete with GINS for Mcm2–7 bind-ing, a function that is relieved in the presence of single-stranded DNA (Bruck et al., 2011; Bruck & Kaplan, 2011,2014, 2015c; Dhingra et al., 2015) (the metazoan Sld2and Sld3 equivalents also bind single-stranded DNA(Bruck & Kaplan, 2015c; Ohlenschlager et al., 2012;Sangrithi et al., 2005)). Thus, GINS binding may not onlybe anticipated by origin melting and the generation ofsingle-stranded DNA, but by strand extrusion as well, asthe separated origin strands may be contained withinthe Mcm2–7 double hexamer for a period of time

(Li et al., 2015; Sun et al., 2014). The topological transfor-mations to DNA in the context of the maturing CMGcomplex are likely aided in by the phosphorylation ofMcm2 by DDK (Bruck & Kaplan, 2009; Lei et al., 1997),an event that weakens the Mcm2/5 gate and promotesring opening (Bruck & Kaplan, 2015a,b). Consistent withthis idea, origin melting and strand extrusion have beenreported to occur concomitantly with DDK-dependentactivation but prior to complete CMG assembly (Bruck &Kaplan, 2015a). Interestingly, recent data suggests thatthe single-stranded DNA-binding activity of buddingyeast Mcm10 is required for initiation and that it mayhelp stabilize the origin melting reaction (Perez-Arnaiz& Kaplan, 2016).

Despite the available insights, the mechanisms thattransition a loaded double hexamer into an active, bi-directional replication fork remain shrouded in mystery.An important future direction will be to understand thetiming of different events in the lifecycle of the helicase,such as when double-hexamer dissociation and repli-some factor recruitment occurs (such as MCM10, Polaand RPA), as compared to origin melting and strandextrusion. The physical mechanisms that coordinateCMG structural changes with origin remodeling likewiseremain ill-defined, and require structural studies of dif-ferent intermediates that, at present, have not been sta-bly isolated.

Mechanism of CMG function

The cracked ring architecture of the MCM2–7 complexis an asset when it comes to loading the helicasearound double-stranded DNA, but would seem to be apotential liability in terms of processive DNA unwindingfunctions (Bochman & Schwacha, 2008; Ilves et al., 2010;Petojevic et al., 2015). Although nucleotide binding byMCM2–7 helps to overcome a natural tendency of thehelicase to splay open (Samel et al., 2014), cracked-ringconformations still readily form (Costa et al., 2011;Lyubimov et al., 2012), providing a physical basis for theweak in vitro helicase activity of MCM2–7 (Bochman &Schwacha, 2008; Ilves et al., 2010). Incorporation ofMCM2–7 into the CMG counteracts the physical defi-ciencies of the MCM2–7 complex for DNA unwinding.3D electron microscopic reconstructions of both theDrosophila and budding yeast CMG have revealed thatGINS and Cdc45 seal off the MCM2/MCM5 discontinuityand reduce the conformational dynamics of the MCM2/5 gate, likely as a means to favor productive ATPaseinterface contacts (Figure 7(A,B)) (Abid Ali et al., 2016;Costa et al., 2011, 2014; Yuan et al., 2016). Thus, insteadof being directly involved in translocation per se, itwould seem that GINS and Cdc45 help shift the

126 M. W. PARKER ET AL.

Page 23: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

structural disposition of the MCM2–7 AAAþ domainsinto productive conformations. Consistent with thisnotion, incorporation of Drosophila MCM2–7 into theCMG 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 (Ilveset al., 2010). Interestingly, in addition to participating inthe control of MCM2–7 ring status, both GINS (Boskovicet al., 2007; Ilves et al., 2010) and Cdc45 (Bruck &Kaplan, 2013; Krastanova et al., 2012; Szambowska et al.,2014) have been shown to bind DNA. Cdc45, which is acatalytically defunct homolog of the RecJ exonuclease(Sanchez-Pulido & Ponting, 2011; Simon et al., 2016),also appears to assist with capture of the leadingstrand during transient CMG gate opening

(Petojevic et al., 2015), thereby possibly playing a pro-tective role during replication fork stalling (Bruck &Kaplan, 2013).

Structural analysis of the CMG reveals the presenceof two distinct conformers that are, as observed for theMCM2–7 helicase core, differentiated by the relativepositioning of the gating subunits (Abid Ali et al., 2016;Yuan et al., 2016). As the MCM NTD is stabilized byextensive contacts with GINS and Cdc45, changeswithin the AAAþ ATPase ring appear primarily respon-sible for the alterations observed between the two con-formational states (Abid Ali et al., 2016; Costa et al.,2011, 2014; Yuan et al., 2016). Altering the position ofthe MCM2/5 gate appears to propagate structuralchanges around the MCM2–7 ring that result in

Figure 7. CMG helicase organization and dynamics. (A) MCM2–7 adopts a spiral, cracked-ring architecture with a discontinuitybetween MCM2 and MCM5. Upon incorporation into the CMG, Cdc45 and GINS convert the helicase into a planar form and sealoff the MCM2/5 gate (EM density maps¼ EMD-1835 and EMD-1833 for MCM2–7 and CMG, respectively). (B) A view of the CMGfrom the AAAþ face illustrating the overall architecture of the complex (PDB¼ 3JC5 (Yuan et al., 2016)). (C) At least two conform-ational states exist for the CMG, a constricted state in which the AAAþ and NTD rings are coplanar (bottom panel), and a relaxedstate where one end of the AAAþ tier lifts up from NTD tier (top panel). These conformations appear coupled to alterations inthe relative disposition of the gating subunits, MCM2 and MCM5 (constricted and relaxed conformer PDB codes¼ 3JC5 and 3JC7,respectively). (D) The activated CMG helicase is thought to translocate along single-stranded DNA, unwinding downstreamtemplate through a combined steric exclusion and DNA wrapping mechanism. A color version of this figure is available at www.tandfonline.com/ibmg.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 127

Page 24: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

significant translational and rotational rearrangement ofMCM6 and, to a lesser extent, MCM4. These changes inturn are coupled to a transition from a constricted con-former state, where the MCM2–7 NTD and CTD tiers areco-planar, to a relaxed, tilted form, which generates anasymmetric expansion of the gap between the NTD andCTD tiers on one side of the ring (Figure 7(C)). It hasbeen suggested that switching between these confor-mations could elicit a pumpjack motion, with the CTDtier rocking with respect to a stable NTD tier duringDNA translocation (Yuan et al., 2016), and that thesedynamics could be used to drive a linear, as opposed torotational, mechanism of DNA translocation (Abid Aliet al., 2016; Yuan et al., 2016). Higher-resolution struc-tures of distinct translocation intermediates, coupledwith single-molecule measurements of ring dynamicsand displacement, will be needed to test these ideas.

An interesting point of discussion in the field hasbeen whether the ATPase regions of MCMs serve as sin-gle or double-stranded DNA translocases. Based on adistant phylogenetic kinship to RuvB, a known double-strand translocase, MCM2–7 was initially proposed tomove along duplex DNA, with the NTD-associatedGINS/Cdc45 accessory subunits serving as a plowshareto separate the two strands (Takahashi et al., 2005).However, subsequent biochemical studies of theDrosophila CMG have shown that the complex requiresa fork for unwinding (Ilves et al., 2010), consistent withboth archaeal MCM and metazoan MCM4,6,7 function-ing as single-stranded DNA translocases on model sub-strates (Ishimi, 1997; Kang et al., 2012; Kelman et al.,1999; Moyer et al., 2006; Shechter et al., 2000). Alongthese lines, studies in archaea have reported that anMCM construct lacking the NTD tier (and that has noassociated GINS/Cdc45) still possesses helicase activity(Barry et al., 2007). Moreover, replisomes have beenchallenged with leading and lagging strand roadblocksin an X. laevis extract system, with only leading strandroadblocks proving capable of stalling fork progression(MCM2–7 is a 30!50 helicase that tracks on the leadingstrand) (Fu et al., 2011). Consistent with this observa-tion, recent analysis of the pattern of forked DNA cross-linking to the CMG reveals that the leading strandshows robust crosslinking to all MCM2–7 subunits, aswell as Cdc45, while lagging strand crosslinks are mark-edly fewer and those that occur are all on the exteriorsurface of MCM2–7 (Petojevic et al., 2015). Collectively,these data strongly indicate that the CMG functions asa single-stranded DNA translocase.

Structural analysis of the apo and substrate-boundCMG states also support a single-strand trackingmodel. In particular, upon binding of the CMG to aforked duplex substrate, nucleic acid density can be

visualized within the central channel whose dimen-sions appear sufficient to accommodate single butnot double-stranded DNA (Abid Ali et al., 2016). Thisfinding is consistent with a previous structural ana-lysis of the CMG bound to duplex DNA containing asingle 30-tail, which showed that the duplex region ofa DNA fork is positioned outside the central channelin an orientation consistent with the known 30!50

polarity of DNA engagement (Costa et al., 2014).Interestingly, the high-resolution structure of the S.cerevisiae CMG reveals that the WH domains of Mcm5and Mcm6, which are disordered within the contextof the double hexamer (Li et al., 2015), becomeordered and take up a position within the centralpore of the CMG that constricts the AAAþ pore to apoint where it is too small to accommodate double-stranded DNA (Yuan et al., 2016). Overall, these linesof evidence largely corroborate the view that theCMG tracks along single-stranded templates, albeitwith a potential capacity to switch to a double-stranded DNA mode of translocation, as might occurduring replication termination, when forks converge(Dewar et al., 2015) (parenthetically, there is evidencethat the E. coli DnaB helicase also can undergo atransition between single and double-stranded trans-location modes (Kaplan, 2000; Kaplan & O'Donnell,2002)). Understanding the role of the WH domainsawaits determination of substrate-bound CMG struc-tures in which the disposition of the nucleic acid sub-strate and the WH domains are fully defined.

Overall, while a variety of mechanistic models existto account for hexameric helicase activity, the so-called steric exclusion model appears most compat-ible with the observed data to date for the CMG. Thesteric exclusion model posits that translocation occursalong single-stranded DNA with the nontemplatestrand being excluded from the central channel. Oneassumption of this framework has been that the lag-ging strand is functionally passive, implying that oncedisplaced, it is inconsequential to helicase function.However, emerging lines of evidence are starting tosuggest that the lagging strand may wrap around theexternal surface of the CMG and support helicaseactivity (Figure 7(D)). Consistent with this notion, DNAinteractions with the external surface of archaealMCM is integral to helicase function (Costa et al.,2008; Graham et al., 2011, 2016; Rothenberg et al.,2007). DNA binding to the MCM external surfaceappears at least partially facilitated by the NTD-Aregion, which can undergo large structural rearrange-ments to expose a DNA-binding site (Chen et al.,2005; Costa et al., 2008; Fletcher et al., 2003; Milleret al., 2014). Whether the eukaryotic MCM NTD-A also

128 M. W. PARKER ET AL.

Page 25: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

facilitates DNA binding is unclear; however, the MCMAAAþ domain contains a conserved, surface-exposedb–hairpin that has been reported to directly engageDNA (Brewster et al., 2010; Graham et al., 2011) and,in the context of the eukaryotic CMG, is integral tohelicase function (Petojevic et al., 2015). Thus, theCMG may function by a mechanism of single-strandedDNA translocation that results in both steric exclusionand the wrapping of lagging strand DNA (Grahamet al., 2011).

Regulatory mechanisms for preventingre-initiation

Origin licensing and firing are the most tightly regu-lated events in the process of DNA replication initi-ation. Multiple interdependent and redundantmechanisms ensure that origins can be marked andutilized only within a certain window of the cell cycle.While the previous discussion of helicase activationhas demonstrated the role of CDK and DDK in posi-tively regulating initiation, a plethora of mechanismsexist to negatively regulate replication licensing andprevent re-replication as well. This section will discusssome general principles for preventing re-replication;although regulatory approaches show substantialevolutionary diversification and expansion, certaincommon themes do exist. For a more thorough dis-cussion of the regulation of DNA replication initiationin eukaryotes, we point the reader to a number ofdedicated reviews on the topic (see (Arias & Walter,2007; Araki, 2010; Hook et al., 2007; Masai et al.,2010; Siddiqui et al., 2013)).

Initiator and helicase regulation

By temporally separating helicase loading from activa-tion, S-CDK can simultaneously activate the helicasewhile also inhibiting helicase-loading factors to preventorigin re-firing. A major means by which CDK protectsagainst re-licensing in S-phase is to regulate initiatorinteractions and their association with chromatin.Although budding yeast ORC remains stably boundthroughout the cell cycle (Diffley et al., 1994; Fujitaet al., 1998; Liang & Stillman, 1997), vertebrate Orc1 isthe target of CDK-dependent and CDK-independentmechanisms that remove it from chromatin (Figure 8)(Findeisen et al., 1999; Kreitz et al., 2001; Li &DePamphilis, 2002; Li et al., 2004; Natale et al., 2000;Rowles et al., 1999; Sun et al., 2002). In both yeastand metazoans, initiation factors are targeted for prote-olysis in a CDK-dependent manner, with buddingyeast exclusively targeting Cdc6 (Drury et al., 1997;

Mimura et al., 2004; Perkins et al., 2001; Piatti et al.,1995) and metazoans targeting both Orc1 and Cdc6(Kalfalah et al., 2015; Lidonnici et al., 2004; Mendezet al., 2002; Ohta et al., 2003; Tatsumi et al., 2003). In S.pombe, Orc2 is also targeted by CDK to prevent re-licensing, but the mechanism utilized in this instance isunclear (Vas et al., 2001; Wuarin et al., 2002). In S. cerevi-siae, CDK-dependent control is further extended toregulate Cdc6 transcription and nuclear localization(Honey & Futcher, 2007; Moll et al., 1991); metazoanCdc6 likewise undergoes CDK-dependent nuclearexport to limit replication (Jiang et al., 1999b; Petersenet al., 1999; Saha et al., 1998). Interestingly, buddingyeast Orc6 and Cdc6 stably associate with CDK duringS-phase and this interaction sterically inhibits interac-tions necessary for pre-RC assembly, such as with Cdt1(Chen & Bell, 2011; Mimura et al., 2004; Wilmes et al.,2004). In metazoans, Cdt1 has not yet been reported tointeract with Orc6.

In addition to regulating ORC and Cdc6, in S.cerevisiae CDK targets Mcm2–7 to prevent new, pro-ductive interactions with chromatin-bound ORC·Cdc6complexes. This action occurs through the CDK-dependent nuclear exclusion and export of Mcm2–7(Labib et al., 1999; Liku et al., 2005; Nguyen et al.,2001; Tanaka & Diffley, 2002). There is currently noevidence that CDK targets metazoan MCM2–7 directly,although phosphorylation of the initiator reduces

Figure 8. Mechanisms to prevent re-replication. Multiple,redundant mechanisms are utilized to prevent re-licensing oforigins after S-phase has initiated. Whereas yeast seems toexclusively utilize CDK-dependent mechanisms to preventre-licensing, metazoans also employ CDK-independent path-ways for negatively regulating Cdt1 activity. A color version ofthis figure is available at www.tandfonline.com/ibmg.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 129

Page 26: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

helicase association with origins (Findeisen et al.,1999). Interestingly, a new, CDK-independent MCMregulatory mechanism has recently been identified inbudding yeast that involves the SUMOylation of allmembers of the Mcm2–7 hexamer in G1. The SUMOmodification in turn, through phosphatase recruit-ment, prevents phosphorylation-dependent helicaseactivation and thus negatively regulates initiation(Wei & Zhao, 2016). How the SUMO pathway integra-tes with the other mechanisms that regulate initiationis currently unclear.

Cdt1: a master regulatory nexus

Cdt1 plays an essential part in the loading of MCM2–7onto DNA as a stable double hexamer. As such, Cdt1turns out to be a common regulatory point.Saccharomyces cerevisiae Cdt1 activity is restricted bythe CDK-dependent inhibition of its interaction withOrc6 (Chen & Bell, 2011) and by nuclear export of Cdt1in G1 (Tanaka & Diffley, 2002). Conversely, metazoanand S. pombe Cdt1 protein levels are regulated in acell-cycle-dependent manner such that Cdt1 isactively degraded upon S-phase entry (Figure 8)(Gopalakrishnan et al., 2001; Nishitani et al., 2000, 2001,2004; Wohlschlegel et al., 2000; Zhong et al., 2003).Degradation of Cdt1 is restricted to S-phase by an inter-action with chromatin-bound PCNA, which serves as aplatform for Cdt1 recognition and ubiquitination by theCullin-RING ligase 4 (CRL4) ubiquitin ligase (Arias &Walter, 2005, 2006; Guarino et al., 2011; Hu & Xiong,2006; Senga et al., 2006). A conserved PCNA-interactingprotein (PIP) degron within the Cdt1 N-terminus facili-tates the interaction with chromatin-bound PCNA and isrequired for degradation (Havens & Walter, 2009, 2011;Senga et al., 2006). Similarly, the CDK-dependent phos-phorylation of Cdt1 leads to its recognition and ubiqui-tination by the SCFSkp2 E3 ubiquitin ligase, providing anadditional mechanism to limit Cdt1 protein levels andprevent re-replication (Kondo et al., 2004; Li et al., 2003;Liu et al., 2004; Nishitani et al., 2006; Sugimoto et al.,2004; Thomer et al., 2004 ). Notably, Cdt1 ubiquitinationcan be reversed through the function of the ubiquitinhydrolase USP37, whose activity stabilizes Cdt1 and pro-motes helicase loading (Hernandez-Perez et al., 2016).

Metazoan Cdt1 is also uniquely regulated by bindingto a partner protein, Geminin. Geminin is a coiled-coilprotein (Lee et al., 2004; Saxena et al., 2004) that wasinitially identified in Xenopus egg extract screens forproteins destabilized in mitosis (McGarry & Kirschner,1998). This study, along with many others, revealed thatGeminin, a nuclear protein whose levels become ele-vated during S-phase, targets and restricts Cdt1 activity

to G1, thus preventing Cdt1-induced re-replication(Cook et al., 2004; Kerns et al., 2007; Lutzmann et al.,2006; Mihaylov et al., 2002; Quinn et al., 2001; Tadaet al., 2001; Yoshida et al., 2005). Geminin binds directlyto Cdt1 (De Marco et al., 2009; Lutzmann et al., 2006;Wohlschlegel et al., 2000), and this interaction has beenshown to inhibit Cdt1 binding to mouse MCM6 (Yanagiet al., 2002). Recent work demonstrates that Gemininbinding to Cdt1 inhibits ORC-dependent helicaseloading but not recruitment (Wu et al., 2014a), afinding consistent with studies in budding yeast show-ing Cdt1-independent recruitment of the helicase tothe origin-bound initiator (Fernandez-Cid et al., 2013;Frigola et al., 2013).

Interestingly, Geminin serves two seemingly dichot-omous functions. Although clearly an inhibitor of licens-ing, Geminin is also required for replication initiationthrough its ability to stabilize Cdt1 levels by protectingthe protein from degradation (Ballabeni et al., 2004;Narasimhachar & Coue, 2009). Thus, metazoans regulateGeminin protein levels and cellular localization to liber-ate Cdt1 from Geminin and promote pre-RC assemblyin G1 (Dimaki et al., 2013; McGarry & Kirschner, 1998;Tsunematsu et al., 2013). In addition, the stoichiometryof the Cdt1·Geminin complex has been suggested toregulate Geminin’s activity, such that a lower orderCdt1:Geminin complex (1:2) is permissive to pre-RCassembly whereas a higher order (2:4) complex is not(De Marco et al., 2009). How this stoichiometry is regu-lated in the context of the other regulatory mechanismsgoverning Geminin and Cdt1 protein levels and cellularlocalization is not known.

Concluding remarks

At this point, many of the major events and inter-mediates that facilitate pre-RC and pre-IC assemblyhave been defined. However, between these stableintermediates there exist multiple, undefined dynamicand transient protein interactions, modifications andexchanges that represent critical steps toward build-ing a replisome. The questions that remain are bothbroad and specific: how does ORC balance sequencepreference and trans-acting chromatin contextualcues when selecting origins? What is the role of Cdt1in MCM2–7 loading and, given the minimal conserva-tion of this protein, is Cdt1 mechanism conservedacross eukarya? Does MCM2–7 melt origins andunwind duplex DNA (as part of the CMG) using anoverlapping or mutually exclusive set of protein-nucleic acid interactions, and how is this functionalswitch regulated? What type of molecular dancemust take place to transition the pre-RC product, the

130 M. W. PARKER ET AL.

Page 27: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

double-stranded DNA-bound MCM2–7double hexamer, into the CMG, and then again intoa bi-directional replication fork defined by two singlehelicases bound to single-stranded DNA? It will becritical to address these and other questions in mul-tiple model eukaryotic organisms to begin to under-stand what level of conservation can be expected forsuch a complex process. Future work that aims toresolve these and other questions will undoubtedlyreveal exciting new mechanisms that underlie thehighly orchestrated and regulated process of replica-tion initiation in eukaryotes.

Acknowledgements

The authors acknowledge Berger and Botchan lab memberspast and present for valuable discussion and helpful advicein preparing this review.

Disclosure statement

The authors report no conflicts of interest. The authors aloneare responsible for the content and writing of this article.

Funding

This work was supported by an NIH NRSA postdoctoral fel-lowship (F32GM116393, to MWP) and by the NCI(R01CA030490, to JMB and MRB).

ORCID

James M. Berger http://orcid.org/0000-0003-0666-1240

References

Abid Ali F, Renault L, Gannon J, et al. (2016). Cryo-EM struc-tures of the eukaryotic replicative helicase bound to atranslocation substrate. Nat Commun 7:10708.

Abreu CM, Kumar R, Hamilton D, et al. (2013). Site-specificphosphorylation of the DNA damage response mediatorrad9 by cyclin-dependent kinases regulates activation ofcheckpoint kinase 1. PLoS Genet 9:e1003310.

Adachi Y, Usukura J, Yanagida M. (1997). A globular complexformation by Nda1 and the other five members of theMCM protein family in fission yeast. Genes Cells 2:467–79.

Aggarwal BD, Calvi BR. (2004). Chromatin regulates originactivity in Drosophila follicle cells. Nature 430:372–6.

Aladjem MI. (2007). Replication in context: dynamic regula-tion of DNA replication patterns in metazoans. Nat RevGenet 8:588–600.

Aparicio OM, Weinstein DM, Bell SP. (1997). Components anddynamics of DNA replication complexes in S. cerevisiae:redistribution of MCM proteins and Cdc45p during Sphase. Cell 91:59–69.

Araki H. (2010). Cyclin-dependent kinase-dependent initiationof chromosomal DNA replication. Curr Opin Cell Biol22:766–71.

Araki H, Leem SH, Phongdara A, Sugino A. (1995). Dpb11,which interacts with DNA polymerase II(epsilon) inSaccharomyces cerevisiae, has a dual role in S-phase pro-gression and at a cell cycle checkpoint. Proc Natl Acad SciUSA92:11791–5.

Aravind L, Landsman D. (1998). AT-hook motifs identified in awide variety of DNA-binding proteins. Nucleic Acids Res26:4413–21.

Arias EE, Walter JC. (2005). Replication-dependent destructionof Cdt1 limits DNA replication to a single round per cellcycle in Xenopus egg extracts. Genes Dev 19:114–26.

Arias EE, Walter JC. (2006). PCNA functions as a molecularplatform to trigger Cdt1 destruction and prevent re-repli-cation. Nat Cell Biol 8:84–90.

Arias EE, Walter JC. (2007). Strength in numbers: preventingrereplication via multiple mechanisms in eukaryotic cells.Genes Dev 21:497–518.

Asano T, Makise M, Takehara M, Mizushima T. (2007).Interaction between ORC and Cdt1p of Saccharomyces cer-evisiae. FEMS Yeast Res 7:1256–62.

Atanasiu C, Deng Z, Wiedmer A, et al. (2006). ORC binding toTRF2 stimulates OriP replication. EMBO Rep 7:716–21.

Aves SJ, Liu Y, Richards TA. (2012). Evolutionary diversificationof eukaryotic DNA replication machinery. Subcell Biochem62:19–35.

Bae B, Chen YH, Costa A, et al. (2009). Insights into the archi-tecture of the replicative helicase from the structure of anarchaeal MCM homolog. Structure 17:211–22.

Balasov M, Huijbregts RP, Chesnokov I. (2007). Role of theOrc6 protein in origin recognition complex-dependentDNA binding and replication in Drosophila melanogaster.Mol Cell Biol 27:3143–53.

Ballabeni A, Melixetian M, Zamponi R, et al. (2004). Humangeminin promotes pre-RC formation and DNA replicationby stabilizing CDT1 in mitosis. EMBO J 23:3122–32.

Barry ER, Bell SD. (2006). DNA replication in the archaea.Microbiol Mol Biol Rev 70:876–87.

Barry ER, Lovett JE, Costa A, et al. (2009). Intersubunit allo-steric communication mediated by a conserved loop inthe MCM helicase. Proc Natl Acad Sci USA106:1051–6.

Barry ER, Mcgeoch AT, Kelman Z, Bell SD. (2007). ArchaealMCM has separable processivity, substrate choice and heli-case domains. Nucleic Acids Res 35:988–98.

Bell SP, Kobayashi R, Stillman B. (1993). Yeast origin recogni-tion complex functions in transcription silencing and DNAreplication. Science 262:1844–9.

Bell SP, Mitchell J, Leber J, et al. (1995). The multidomainstructure of Orc1p reveals similarity to regulators of DNAreplication and transcriptional silencing. Cell 83:563–8.

Bell SP, Stillman B. (1992). ATP-dependent recognition ofeukaryotic origins of DNA replication by a multiproteincomplex. Nature 357:128–34.

Berbenetz NM, Nislow C, Brown GW. (2010). Diversity ofeukaryotic DNA replication origins revealed by genome-wide analysis of chromatin structure. PLoS Genet6:e1001092.

Bianconi E, Piovesan A, Facchin F, et al. (2013). An estimationof the number of cells in the human body. Ann Hum Biol40:463–71.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 131

Page 28: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Bleichert F, Balasov M, Chesnokov I, et al. (2013). A Meier-Gorlin syndrome mutation in a conserved C-terminal helixof Orc6 impedes origin recognition complex formation.Elife 2:e00882.

Bleichert F, Botchan MR, Berger JM. (2015). Crystal structureof the eukaryotic origin recognition complex. Nature519:321–6.

Blow JJ. (1993). Preventing re-replication of DNA in a singlecell cycle: evidence for a replication licensing factor. J CellBiol 122:993–1002.

Blow JJ, Laskey RA. (1988). A role for the nuclear envelope incontrolling DNA replication within the cell cycle. Nature332:546–8.

Bochman ML, Bell SP, Schwacha A. (2008). Subunit organiza-tion of Mcm2-7 and the unequal role of active sites in ATPhydrolysis and viability. Mol Cell Biol 28:5865–73.

Bochman ML, Schwacha A. (2008). The Mcm2-7 complex hasin vitro helicase activity. Mol Cell 31:287–93.

Bochman ML, Schwacha A. (2015). DNA replication: strandseparation unravelled. Nature 524:166–7.

Boos D, Yekezare M, Diffley JF. (2013). Identification of a het-eromeric complex that promotes DNA replication originfiring in human cells. Science 340:981–4.

Bosco G, Du W, Orr-Weaver TL. (2001). DNA replication con-trol through interaction of E2F-RB and the origin recogni-tion complex. Nat Cell Biol 3:289–95.

Bose ME, Mcconnell KH, Gardner-Aukema KA, et al. (2004).The origin recognition complex and Sir4 protein recruitSir1p to yeast silent chromatin through independent inter-actions requiring a common Sir1p domain. Mol Cell Biol24:774–86.

Boskovic J, Bragado-Nilsson E, Saligram Prabhakar B, et al.(2016). Molecular architecture of the recombinant humanMCM2-7 helicase in complex with nucleotides and DNA.Cell Cycle 15:2431–40.

Boskovic J, Coloma J, Aparicio T, et al. (2007). Moleculararchitecture of the human GINS complex. EMBO Rep8:678–84.

Bowers JL, Randell JC, Chen S, Bell SP. (2004). ATP hydrolysisby ORC catalyzes reiterative Mcm2-7 assembly at a definedorigin of replication. Mol Cell 16:967–78.

Brewer BJ, Fangman WL. (1987). The localization of replica-tion origins on ARS plasmids in S. cerevisiae. Cell51:463–71.

Brewster AS, Slaymaker IM, Afif SA, Chen XS. (2010).Mutational analysis of an archaeal minichromosome main-tenance protein exterior hairpin reveals critical residues forhelicase activity and DNA binding. BMC Mol Biol 11:62.

Brewster AS, Wang G, Yu X, et al. (2008). Crystal structure ofa near-full-length archaeal MCM: functional insights for anAAAþ hexameric helicase. Proc Natl Acad SciUSA105:20191–6.

Broek D, Bartlett R, Crawford K, Nurse P. (1991). Involvementof p34cdc2 in establishing the dependency of S phase onmitosis. Nature 349:388–93.

Bruck I, Kanter DM, Kaplan DL. (2011). Enabling association ofthe GINS protein tetramer with the mini chromosomemaintenance (Mcm)2-7 protein complex by phosphory-lated Sld2 protein and single-stranded origin DNA. J BiolChem 286:36414–26.

Bruck I, Kaplan D. (2009). Dbf4-Cdc7 phosphorylation of Mcm2is required for cell growth. J Biol Chem 284:28823–31.

Bruck I, Kaplan DL. (2011). Origin single-stranded DNAreleases Sld3 protein from the Mcm2-7 complex, allowingthe GINS tetramer to bind the Mcm2-7 complex. J BiolChem 286:18602–13.

Bruck I, Kaplan DL. (2013). Cdc45 protein-single-strandedDNA interaction is important for stalling the helicase dur-ing replication stress. J Biol Chem 288:7550–63.

Bruck I, Kaplan DL. (2014). The replication initiation proteinSld2 regulates helicase assembly. J Biol Chem 289:1948–59.

Bruck I, Kaplan DL. (2015a). Conserved mechanism for coordi-nating replication fork helicase assembly with phosphoryl-ation of the helicase. Proc Natl Acad Sci USA 112:11223–8.

Bruck I, Kaplan DL. (2015b). The Dbf4-Cdc7 kinase promotesMcm2-7 ring opening to allow for single-stranded DNAextrusion and helicase assembly. J Biol Chem 290:1210–21.

Bruck I, Kaplan DL. (2015c). The Replication Initiation ProteinSld3/Treslin Orchestrates the Assembly of the ReplicationFork Helicase during S Phase. J Biol Chem 290:27414–24.

Buchman AR, Kimmerly WJ, Rine J, Kornberg RD. (1988). TwoDNA-binding factors recognize specific sequences atsilencers, upstream activating sequences, autonomouslyreplicating sequences, and telomeres in Saccharomycescerevisiae. Mol Cell Biol 8:210–25.

Burkhart R, Schulte D, Hu D, et al. (1995). Interactions ofhuman nuclear proteins P1Mcm3 and P1Cdc46. Eur JBiochem 228:431–8.

Cadoret JC, Meisch F, Hassan-Zadeh V, et al. (2008). Genome-wide studies highlight indirect links between human repli-cation origins and gene regulation. Proc Natl Acad Sci USA105:15837–42.

Calzada A, Hodgson B, Kanemaki M, et al. (2005). Molecularanatomy and regulation of a stable replisome at a pausedeukaryotic DNA replication fork. Genes Dev 19:1905–19.

Carpenter PB, Mueller PR, Dunphy WG. (1996). Role for aXenopus Orc2-related protein in controlling DNA replica-tion. Nature 379:357–60.

Cayrou C, Ballester B, Peiffer I, et al. (2015). The chromatinenvironment shapes DNA replication origin organizationand defines origin classes. Genome Res 25:1873–85.

Cayrou C, Coulombe P, Puy A, et al. (2012). New insights intoreplication origin characteristics in metazoans. Cell Cycle11:658–67.

Cayrou C, Coulombe P, Vigneron A, et al. (2011). Genome-scale analysis of metazoan replication origins reveals theirorganization in specific but flexible sites defined by con-served features. Genome Res 21:1438–49.

Chakraborty A, Shen Z, Prasanth SG. (2011). “Initiation tochromatin organization” on heterochromatin: linking DNAreplication initiation to chromatin. Epigenetics 6:665–70.

Chang F, Riera A, Evrin C, et al. (2015). Cdc6 ATPase activitydisengages Cdc6 from the pre-replicative complex topromote DNA replication. Elife 4:e05795. doi: 10.7554/eLife.05795

Chapman JW, Johnston LH. (1989). The yeast gene, DBF4,essential for entry into S phase is cell cycle regulated. ExpCell Res 180:419–28.

Chen S, Bell SP. (2011). CDK prevents Mcm2-7 helicase load-ing by inhibiting Cdt1 interaction with Orc6. Genes Dev25:363–72.

Chen S, De Vries MA, Bell SP. (2007). Orc6 is required fordynamic recruitment of Cdt1 during repeated Mcm2-7loading. Genes Dev 21:2897–907.

132 M. W. PARKER ET AL.

Page 29: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Chen Y, Hennessy KM, Botstein D, Tye BK. (1992). CDC46/MCM5, a yeast protein whose subcellular localization iscell cycle-regulated, is involved in DNA replication atautonomously replicating sequences. Proc Natl Acad SciUSA 89:10459–63.

Chen YJ, Yu X, Kasiviswanathan R, et al. (2005). Structuralpolymorphism of Methanothermobacter thermautotrophi-cus MCM. J Mol Biol 346:389–94.

Chen Z, Speck C, Wendel P, et al. (2008). The architecture ofthe DNA replication origin recognition complex inSaccharomyces cerevisiae. Proc Natl Acad Sci USA105:10326–31.

Cheng L, Collyer T, Hardy CF. (1999). Cell cycle regulation ofDNA replication initiator factor Dbf4p. Mol Cell Biol19:4270–8.

Chesnokov I, Gossen M, Remus D, Botchan M. (1999).Assembly of functionally active Drosophila origin recogni-tion complex from recombinant proteins. Genes Dev13:1289–96.

Chesnokov I, Remus D, Botchan M. (2001). Functional analysisof mutant and wild-type Drosophila origin recognitioncomplex. Proc Natl Acad Sci USA 98:11997–2002.

Chesnokov IN. (2007). Multiple functions of the origin recog-nition complex. Int Rev Cytol 256:69–109.

Chesnokov IN, Chesnokova ON, Botchan M. (2003). A cytoki-netic function of Drosophila ORC6 protein resides in adomain distinct from its replication activity. Proc Natl AcadSci USA 100:9150–5.

Cho WH, Lee YJ, Kong SI, et al. (2006). CDC7 kinase phos-phorylates serine residues adjacent to acidic amino acidsin the minichromosome maintenance 2 protein. Proc NatlAcad Sci USA 103:11521–6.

Chong JP, Hayashi MK, Simon MN, et al. (2000). A double-hexamer archaeal minichromosome maintenance proteinis an ATP-dependent DNA helicase. Proc Natl Acad Sci USA97:1530–5.

Chong JP, Mahbubani HM, Khoo CY, Blow JJ. (1995).Purification of an MCM-containing complex as a compo-nent of the DNA replication licensing system. Nature375:418–21.

Chong JP, Thommes P, Blow JJ. (1996). The role of MCM/P1proteins in the licensing of DNA replication. TrendsBiochem Sci 21:102–6.

Chuang RY, Chretien L, Dai J, Kelly TJ. (2002). Purification andcharacterization of the Schizosaccharomyces pombe originrecognition complex: interaction with origin DNA andCdc18 protein. J Biol Chem 277:16920–7.

Chuang RY, Kelly TJ. (1999). The fission yeast homologue ofOrc4p binds to replication origin DNA via multiple AT-hooks. Proc Natl Acad Sci USA 96:2656–61.

Clarey MG, Botchan M, Nogales E. (2008). Single particle EMstudies of the Drosophila melanogaster origin recognitioncomplex and evidence for DNA wrapping. J Struct Biol164:241–9.

Clarey MG, Erzberger JP, Grob P, et al. (2006). Nucleotide-dependent conformational changes in the DnaA-like coreof the origin recognition complex. Nat Struct Mol Biol13:684–90.

Claycomb JM, Macalpine DM, Evans JG, et al. (2002).Visualization of replication initiation and elongation inDrosophila. J Cell Biol 159:225–36.

Clyne RK, Kelly TJ. (1995). Genetic analysis of an ARS elementfrom the fission yeast Schizosaccharomyces pombe. EMBOJ 14:6348–57.

Cocker JH, Piatti S, Santocanale C, et al. (1996). An essentialrole for the Cdc6 protein in forming the pre-replicativecomplexes of budding yeast. Nature 379:180–2.

Coleman TR, Carpenter PB, Dunphy WG. (1996). The XenopusCdc6 protein is essential for the initiation of a single roundof DNA replication in cell-free extracts. Cell 87:53–63.

Cook JG, Chasse DA, Nevins JR. (2004). The regulated associ-ation of Cdt1 with minichromosome maintenance proteinsand Cdc6 in mammalian cells. J Biol Chem 279:9625–33.

Costa A, Hood IV, Berger JM. (2013). Mechanisms for initiat-ing cellular DNA replication. Annu Rev Biochem 82:25–54.

Costa A, Ilves I, Tamberg N, et al. (2011). The structural basisfor MCM2-7 helicase activation by GINS and Cdc45. NatStruct Mol Biol 18:471–7.

Costa A, Renault L, Swuec P, et al. (2014). DNA binding polar-ity, dimerization, and ATPase ring remodeling in the CMGhelicase of the eukaryotic replisome. Elife 3:e03273.

Costa A, Van Duinen G, Medagli B, et al. (2008). Cryo-electronmicroscopy reveals a novel DNA-binding site on the MCMhelicase. EMBO J 27:2250–8.

Costas C, De La Paz Sanchez M, Stroud H, et al. (2011).Genome-wide mapping of Arabidopsis thaliana origins ofDNA replication and their associated epigenetic marks. NatStruct Mol Biol 18:395–400.

Coster G, Frigola J, Beuron F, et al. (2014). Origin licensingrequires ATP binding and hydrolysis by the MCM replica-tive helicase. Mol Cell 55:666–77.

Creager RL, Li Y, Macalpine DM. (2015). SnapShot: Origins ofDNA replication. Cell 161:418e1.

Crevel G, Ivetic A, Ohno K, et al. (2001). Nearest neighbouranalysis of MCM protein complexes in Drosophila mela-nogaster. Nucleic Acids Res 29:4834–42.

Dai J, Chuang RY, Kelly TJ. (2005). DNA replication origins inthe Schizosaccharomyces pombe genome. Proc Natl AcadSci USA 102:337–42.

Dalton S, Hopwood B. (1997). Characterization of Cdc47p-minichromosome maintenance complexes inSaccharomyces cerevisiae: identification of Cdc45p as asubunit. Mol Cell Biol 17:5867–75.

Dalton S, Whitbread L. (1995). Cell cycle-regulated nuclearimport and export of Cdc47, a protein essential for initi-ation of DNA replication in budding yeast. Proc Natl AcadSci USA 92:2514–18.

Davey MJ, Indiani C, O'donnell M. (2003). Reconstitution ofthe Mcm2-7p heterohexamer, subunit arrangement, andATP site architecture. J Biol Chem 278:4491–9.

De Marco V, Gillespie PJ, Li A, et al. (2009). Quaternary struc-ture of the human Cdt1-Geminin complex regulates DNAreplication licensing. Proc Natl Acad Sci USA106:19807–12.

Deegan TD, Yeeles JT, Diffley JF. (2016). Phosphopeptidebinding by Sld3 links Dbf4-dependent kinase to MCM rep-licative helicase activation. EMBO J 35:961–73.

Delgado S, Gomez M, Bird A, Antequera F. (1998). Initiationof DNA replication at CpG islands in mammalian chromo-somes. EMBO J 17:2426–35.

Deng Z, Norseen J, Wiedmer A, et al. (2009). TERRA RNAbinding to TRF2 facilitates heterochromatin formation andORC recruitment at telomeres. Mol Cell 35:403–13.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 133

Page 30: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Devault A, Gueydon E, Schwob E. (2008). Interplaybetween S-cyclin-dependent kinase and Dbf4-dependentkinase in controlling DNA replication through phosphoryl-ation of yeast Mcm4 N-terminal domain. Mol Biol Cell19:2267–77.

Devault A, Vallen EA, Yuan T, et al. (2002). Identificationof Tah11/Sid2 as the ortholog of the replication licensingfactor Cdt1 in Saccharomyces cerevisiae. Curr Biol12:689–94.

Dewar JM, Budzowska M, Walter JC. (2015). The mechanismof DNA replication termination in vertebrates. Nature525:345–50.

Dhar SK, Delmolino L, Dutta A. (2001). Architecture of thehuman origin recognition complex. J Biol Chem276:29067–71.

Dhar SK, Dutta A. (2000). Identification and characterizationof the human ORC6 homolog. J Biol Chem 275:34983–8.

Dhingra N, Bruck I, Smith S, et al. (2015). Dpb11 protein helpscontrol assembly of the Cdc45�Mcm2-7�GINS replicationfork helicase . J Biol Chem 290:7586–601.

Diffley JF, Cocker JH. (1992). Protein-DNA interactions at ayeast replication origin. Nature 357:169–72.

Diffley JF, Cocker JH, Dowell SJ, Rowley A. (1994). Two stepsin the assembly of complexes at yeast replication originsin vivo. Cell 78:303–16.

Diffley JF, Stillman B. (1988). Purification of a yeast proteinthat binds to origins of DNA replication and a transcrip-tional silencer. Proc Natl Acad Sci USA 85:2120–4.

Dimaki M, Xouri G, Symeonidou IE, et al. (2013). Cell cycle-dependent subcellular translocation of the human DNAlicensing inhibitor geminin. J Biol Chem 288:23953–63.

Donovan JD, Toyn JH, Johnson AL, Johnston LH. (1994).P40SDB25, a putative CDK inhibitor, has a role in theM/G1 transition in Saccharomyces cerevisiae. Genes Dev8:1640–53.

Donovan S, Harwood J, Drury LS, Diffley JF. (1997). Cdc6p-dependent loading of Mcm proteins onto pre-replicativechromatin in budding yeast. Proc Natl Acad Sci USA94:5611–16.

Douglas ME, Diffley JF. (2016). Recruitment of Mcm10 toSites of Replication Initiation Requires Direct Binding tothe Minichromosome Maintenance (MCM) Complex. J BiolChem 291:5879–88.

Dowell SJ, Romanowski P, Diffley JF. (1994). Interaction ofDbf4, the Cdc7 protein kinase regulatory subunit, withyeast replication origins in vivo. Science 265:1243–6.

Drury LS, Perkins G, Diffley JF. (1997). The Cdc4/34/53 path-way targets Cdc6p for proteolysis in budding yeast. EMBOJ 16:5966–76.

Dueber EC, Costa A, Corn JE, et al. (2011). Molecular determi-nants of origin discrimination by Orc1 initiators in archaea.Nucleic Acids Res 39:3621–31.

Dueber EL, Corn JE, Bell SD, Berger JM. (2007). Replicationorigin recognition and deformation by a heterodimericarchaeal Orc1 complex. Science 317:1210–13.

Duzdevich D, Warner MD, Ticau S, et al. (2015). The dynamicsof eukaryotic replication initiation: origin specificity, licens-ing, and firing at the single-molecule level. Mol Cell58:483–94.

Eaton ML, Galani K, Kang S, et al. (2010). Conserved nucleo-some positioning defines replication origins. Genes Dev24:748–53.

Eaton ML, Prinz JA, Macalpine HK, et al. (2011). Chromatinsignatures of the Drosophila replication program. GenomeRes 21:164–74.

Edgell DR, Doolittle WF. (1997). Archaea and the origin(s) ofDNA replication proteins. Cell 89:995–8.

Edwards MC, Tutter AV, Cvetic C, et al. (2002). MCM2-7 com-plexes bind chromatin in a distributed pattern surroundingthe origin recognition complex in Xenopus egg extracts.J Biol Chem 277:33049–57.

El-Sayed NM, Myler PJ, Bartholomeu DC, et al. (2005). Thegenome sequence of Trypanosoma cruzi, etiologic agentof Chagas disease. Science 309:409–15.

Enemark EJ, Joshua-Tor L. (2006). Mechanism of DNA trans-location in a replicative hexameric helicase. Nature442:270–5.

Enemark EJ, Joshua-Tor L. (2008). On helicases and othermotor proteins. Curr Opin Struct Biol 18:243–57.

Erzberger JP, Berger JM. (2006). Evolutionary relationshipsand structural mechanisms of AAAþ proteins. Annu RevBiophys Biomol Struct 35:93–114.

Evrin C, Clarke P, Zech J, et al. (2009). A double-hexamericMCM2-7 complex is loaded onto origin DNA during licens-ing of eukaryotic DNA replication. Proc Natl Acad Sci USA106:20240–5.

Evrin C, Fernandez-Cid A, Riera A, et al. (2014). The ORC/Cdc6/MCM2-7 complex facilitates MCM2-7 dimerizationduring prereplicative complex formation. Nucleic Acids Res42:2257–69.

Evrin C, Fernandez-Cid A, Zech J, et al. (2013). In the absenceof ATPase activity, pre-RC formation is blocked prior toMCM2-7 hexamer dimerization. Nucleic Acids Res41:3162–72.

Fang D, Cao Q, Lou H. (2016). Sld3-MCM Interaction Facilitatedby Dbf4-Dependent Kinase Defines an Essential Step inEukaryotic DNA Replication Initiation. Front Microbiol 7:885.

Feng L, Hu Y, Wang B, et al. (2000). Loss control of Mcm5interaction with chromatin in cdc6-1 mutated in CDC-NTPmotif. DNA Cell Biol 19:447–57.

Fenouil R, Cauchy P, Koch F, et al. (2012). CpG islands andGC content dictate nucleosome depletion in a transcrip-tion-independent manner at mammalian promoters.Genome Res 22:2399–408.

Ferenbach A, Li A, Brito-Martins M, Blow JJ. (2005).Functional domains of the Xenopus replication licensingfactor Cdt1. Nucleic Acids Res 33:316–24.

Fernandez-Cid A, Riera A, Tognetti S, et al. (2013). An ORC/Cdc6/MCM2-7 complex is formed in a multistep reactionto serve as a platform for MCM double-hexamer assembly.Mol Cell 50:577–88.

Findeisen M, El-Denary M, Kapitza T, et al. (1999). CyclinA-dependent kinase activity affects chromatin binding ofORC, Cdc6, and MCM in egg extracts of Xenopus laevis.Eur J Biochem 264:415–26.

Fiorani P, Bjornsti MA. (2000). Mechanisms of DNA topoisom-erase I-induced cell killing in the yeast Saccharomyces cere-visiae. Ann N Y Acad Sci 922:65–75.

Fletcher RJ, Bishop BE, Leon RP, et al. (2003). Thestructure and function of MCM from archaeal M.Thermoautotrophicum. Nat Struct Biol 10:160–7.

Fletcher RJ, Shen J, Gomez-Llorente Y, et al. (2005).Double hexamer disruption and biochemical activities of

134 M. W. PARKER ET AL.

Page 31: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Methanobacterium thermoautotrophicum MCM. J Biol Chem280:42405–10.

Fletcher RJ, Shen J, Holden LG, Chen XS. (2008). Identificationof amino acids important for the biochemical activityof Methanothermobacter thermautotrophicus MCM.Biochemistry 47:9981–6.

Foss M, Mcnally FJ, Laurenson P, Rine J. (1993). Origin recog-nition complex (ORC) in transcriptional silencing and DNAreplication in S. cerevisiae. Science 262:1838–44.

Fox CA, Ehrenhofer-Murray AE, Loo S, Rine J. (1997). Theorigin recognition complex, SIR1, and the S phase require-ment for silencing. Science 276:1547–51.

Fox CA, Loo S, Dillin A, Rine J. (1995). The origin recognitioncomplex has essential functions in transcriptional silencingand chromosomal replication. Genes Dev 9:911–24.

Francis LI, Randell JC, Takara TJ, et al. (2009). Incorporationinto the prereplicative complex activates the Mcm2-7 heli-case for Cdc7-Dbf4 phosphorylation. Genes Dev 23:643–54.

Francois Jacob SB, Cuzin F. (1963). On the Regulation of DNAReplication in Bacteria. Cold Spring Harb Symp Quant Biol28:329–48.

Frigola J, Remus D, Mehanna A, Diffley JF. (2013). ATPase-dependent quality control of DNA replication origin licens-ing. Nature 495:339–43.

Froelich CA, Kang S, Epling LB, et al. (2014). A conservedMCM single-stranded DNA binding element is essential forreplication initiation. Elife 3:e01993.

Fu Y, Slaymaker IM, Wang J, et al. (2014). The 1.8-Å crystalstructure of the N-terminal domain of an archaeal MCM asa right-handed filament. J Mol Biol 426:1512–23.

Fu YV, Yardimci H, Long DT, et al. (2011). Selective bypass ofa lagging strand roadblock by the eukaryotic replicativeDNA helicase. Cell 146:931–41.

Fujita M, Hori Y, Shirahige K, et al. (1998). Cell cycle depend-ent topological changes of chromosomal replicationorigins in Saccharomyces cerevisiae. Genes Cells 3:737–49.

Gaczynska M, Osmulski PA, Jiang Y, et al. (2004). Atomicforce microscopic analysis of the binding of theSchizosaccharomyces pombe origin recognition complexand the spOrc4 protein with origin DNA. Proc Natl AcadSci USA 101:17952–7.

Gambus A, Jones RC, Sanchez-Diaz A, et al. (2006). GINSmaintains association of Cdc45 with MCM in replisomeprogression complexes at eukaryotic DNA replication forks.Nat Cell Biol 8:358–66.

Gambus A, Khoudoli GA, Jones RC, Blow JJ. (2011). MCM2-7form double hexamers at licensed origins in Xenopus eggextract. J Biol Chem 286:11855–64.

Gardner KA, Rine J, Fox CA. (1999). A region of the Sir1 pro-tein dedicated to recognition of a silencer and requiredfor interaction with the Orc1 protein in Saccharomyces cer-evisiae. Genetics 151:31–44.

Gaudier M, Schuwirth BS, Westcott SL, Wigley DB. (2007).Structural basis of DNA replication origin recognition byan ORC protein. Science 317:1213–16.

Gavin KA, Hidaka M, Stillman B. (1995). Conserved initiatorproteins in eukaryotes. Science 270:1667–71.

Gibson SI, Surosky RT, Tye BK. (1990). The phenotype of theminichromosome maintenance mutant mcm3 is character-istic of mutants defective in DNA replication. Mol Cell Biol10:5707–20.

Gilbert DM. (2004). In search of the holy replicator. Nat RevMol Cell Biol 5:848–55.

Gillespie PJ, Li A, Blow JJ. (2001). Reconstitution of licensedreplication origins on Xenopus sperm nuclei using purifiedproteins. BMC Biochem 2:15.

Giordano-Coltart J, Ying CY, Gautier J, Hurwitz J. (2005).Studies of the properties of human origin recognitioncomplex and its Walker A motif mutants. Proc Natl AcadSci USA 102:69–74.

Givens RM, Lai WK, Rizzo JM, et al. (2012). Chromatinarchitectures at fission yeast transcriptional promoters andreplication origins. Nucleic Acids Res 40:7176–89.

Gomez-Llorente Y, Fletcher RJ, Chen XS, et al. (2005).Polymorphism and double hexamer structure in the arch-aeal minichromosome maintenance (MCM) helicase fromMethanobacterium thermoautotrophicum. J Biol Chem280:40909–15.

Gopalakrishnan V, Simancek P, Houchens C, et al. (2001).Redundant control of rereplication in fission yeast. ProcNatl Acad Sci USA 98:13114–19.

Gossen M, Pak DT, Hansen SK, et al. (1995). A Drosophilahomolog of the yeast origin recognition complex. Science270:1674–7.

Graham BW, Schauer GD, Leuba SH, Trakselis MA. (2011).Steric exclusion and wrapping of the excluded DNA strandoccurs along discrete external binding paths during MCMhelicase unwinding. Nucleic Acids Res 39:6585–95.

Graham BW, Tao Y, Dodge KL, et al. (2016). DNA interactionsprobed by H/D exchange FT-ICR mass spectrometry con-firm external binding sites on the MCM helicase. J BiolChem 291:12467–80.

Grallert B, Nurse P. (1996). The ORC1 homolog orp1 in fissionyeast plays a key role in regulating onset of S phase.Genes Dev 10:2644–54.

Gregan J, Lindner K, Brimage L, et al. (2003). Fission yeastCdc23/Mcm10 functions after pre-replicative complex for-mation to promote Cdc45 chromatin binding. Mol Biol Cell14:3876–87.

Gros J, Kumar C, Lynch G, et al. (2015). Post-licensingSpecification of Eukaryotic Replication Origins byFacilitated Mcm2-7 Sliding along DNA. Molecular Cell60:797–807.

Guarino E, Shepherd ME, Salguero I, et al. (2011). Cdt1proteolysis is promoted by dual PIP degrons and is modu-lated by PCNA ubiquitylation. Nucleic Acids Res39:5978–90.

Guenther B, Onrust R, Sali A, et al. (1997). Crystal structure ofthe delta' subunit of the clamp-loader complex of E. coliDNA polymerase III. Cell 91:335–45.

Hanson PI, Whiteheart SW. (2005). AAAþ proteins: haveengine, will work. Nat Rev Mol Cell Biol 6:519–29.

Hardy CF, Dryga O, Seematter S, et al. (1997). mcm5/cdc46-bob1 bypasses the requirement for the S phase activatorCdc7p. Proc Natl Acad Sci USA 94:3151–5.

Hartwell LH. (1976). Sequential function of gene productsrelative to DNA synthesis in the yeast cell cycle. J Mol Biol104:803–17.

Harvey KJ, Newport J. (2003). Metazoan origin selection: ori-gin recognition complex chromatin binding is regulatedby CDC6 recruitment and ATP hydrolysis. J Biol Chem278:48524–8.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 135

Page 32: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Havens CG, Walter JC. (2009). Docking of a specialized PIPBox onto chromatin-bound PCNA creates a degron for theubiquitin ligase CRL4Cdt2. Mol Cell 35:93–104.

Havens CG, Walter JC. (2011). Mechanism of CRL4(Cdt2), aPCNA-dependent E3 ubiquitin ligase. Genes Dev25:1568–82.

Hayles J, Fisher D, Woollard A, Nurse P. (1994). Temporalorder of S phase and mitosis in fission yeast is determinedby the state of the p34cdc2-mitotic B cyclin complex. Cell78:813–22.

Heichinger C, Penkett CJ, Bahler J, Nurse P. (2006). Genome-wide characterization of fission yeast DNA replication ori-gins. EMBO J 25:5171–9.

Heinzel SS, Krysan PJ, Tran CT, Calos MP. (1991). AutonomousDNA replication in human cells is affected by the size andthe source of the DNA. Mol Cell Biol 11:2263–72.

Heller RC, Kang S, Lam WM, et al. (2011). Eukaryotic origin-dependent DNA replication in vitro reveals sequentialaction of DDK and S-CDK kinases. Cell 146:80–91.

Hennessy KM, Clark CD, Botstein D. (1990). Subcellular local-ization of yeast CDC46 varies with the cell cycle. GenesDev 4:2252–63.

Hernandez-Perez S, Cabrera E, Amoedo H, et al. (2016).USP37 deubiquitinates Cdt1 and contributes to regulateDNA replication. Mol Oncol 10:1196–206.

Higa M, Kushiyama T, Kurashige S, et al. (2016). TRF2 recruitsORC through TRFH domain dimerization. Biochim BiophysActa 1864:191–201.

Hizume K, Yagura M, Araki H. (2013). Concerted interactionbetween origin recognition complex (ORC), nucleosomesand replication origin DNA ensures stable ORC-origin bind-ing. Genes Cells 18:764–79.

Hoang ML, Leon RP, Pessoa-Brandao L, et al. (2007).Structural changes in Mcm5 protein bypass Cdc7-Dbf4function and reduce replication origin efficiency inSaccharomyces cerevisiae. Mol Cell Biol 27:7594–602.

Hofmann JF, Beach D. (1994). cdt1 is an essential target ofthe Cdc10/Sct1 transcription factor: requirement for DNAreplication and inhibition of mitosis. EMBO J 13:425–34.

Hogan E, Koshland D. (1992). Addition of extra origins of rep-lication to a minichromosome suppresses its mitotic lossin cdc6 and cdc14 mutants of Saccharomyces cerevisiae.Proc Natl Acad Sci USA 89:3098–102.

Hoggard T, Shor E, Muller CA, et al. (2013). A Link betweenORC-origin binding mechanisms and origin activation timerevealed in budding yeast. PLoS Genet 9:e1003798.

Hollingsworth RE, Jr., Sclafani RA. (1990). DNA metabolismgene CDC7 from yeast encodes a serine (threonine) pro-tein kinase. Proc Natl Acad Sci USA 87:6272–6.

Homesley L, Lei M, Kawasaki Y, et al. (2000). Mcm10 and theMCM2-7 complex interact to initiate DNA synthesis and torelease replication factors from origins. Genes Dev14:913–26.

Honey S, Futcher B. (2007). Roles of the CDK phosphorylationsites of yeast Cdc6 in chromatin binding and rereplication.Mol Biol Cell 18:1324–36.

Hook SS, Lin JJ, Dutta A. (2007). Mechanisms to control rere-plication and implications for cancer. Curr Opin Cell Biol19:663–71.

Hopwood B, Dalton S. (1996). Cdc45p assembles into a com-plex with Cdc46p/Mcm5p, is required for minichromosome

maintenance, and is essential for chromosomal DNA repli-cation. Proc Natl Acad Sci USA 93:12309–14.

Hoshina S, Yura K, Teranishi H, et al. (2013). Human originrecognition complex binds preferentially to G-quadruplex-preferable RNA and single-stranded DNA. J Biol Chem288:30161–71.

Houchens CR, Lu W, Chuang RY, et al. (2008). Multiple mech-anisms contribute to Schizosaccharomyces pombe originrecognition complex-DNA interactions. J Biol Chem283:30216–24.

Hu J, Xiong Y. (2006). An evolutionarily conserved function ofproliferating cell nuclear antigen for Cdt1 degradation bythe Cul4-Ddb1 ubiquitin ligase in response to DNA dam-age. J Biol Chem 281:3753–6.

Hua XH, Newport J. (1998). Identification of a preinitiationstep in DNA replication that is independent of origin rec-ognition complex and cdc6, but dependent on cdk2. J CellBiol 140:271–81.

Huang DW, Fanti L, Pak DT, et al. (1998). Distinct cytoplasmicand nuclear fractions of Drosophila heterochromatinprotein 1: their phosphorylation levels and associationswith origin recognition complex proteins. J Cell Biol142:307–18.

Huberman JA, Spotila LD, Nawotka KA, et al. (1987). The invivo replication origin of the yeast 2 microns plasmid. Cell51:473–81.

Huberman JA, Zhu JG, Davis LR, Newlon CS. (1988). Closeassociation of a DNA replication origin and an ARS elem-ent on chromosome III of the yeast, Saccharomyces cerevi-siae. Nucleic Acids Res 16:6373–84.

Huppert JL, Balasubramanian S. (2007). G-quadruplexes inpromoters throughout the human genome. Nucleic AcidsRes 35:406–13.

Hyrien O, Mechali M. (1993). Chromosomal replication ini-tiates and terminates at random sequences but at regularintervals in the ribosomal DNA of Xenopus early embryos.EMBO J 12:4511–20.

Ilves I, Petojevic T, Pesavento JJ, Botchan MR. (2010).Activation of the MCM2-7 helicase by association withCdc45 and GINS proteins. Mol Cell 37:247–58.

Ishimi Y. (1997). A DNA helicase activity is associated with anMCM4, -6, and -7 protein complex. J Biol Chem272:24508–13.

Iyer LM, Leipe DD, Koonin EV, Aravind L. (2004). Evolutionaryhistory and higher order classification of AAAþ ATPases. JStruct Biol 146:11–31.

Jackson AL, Pahl PM, Harrison K, et al. (1993). Cell cycle regu-lation of the yeast Cdc7 protein kinase by association withthe Dbf4 protein. Mol Cell Biol 13:2899–908.

Jares P, Blow JJ. (2000). Xenopus cdc7 function is dependenton licensing but not on XORC, XCdc6, or CDK activity andis required for XCdc45 loading. Genes Dev 14:1528–40.

Jazwinski SM, Edelman GM. (1982). Protein complexes fromactive replicative fractions associate in vitro with the repli-cation origins of yeast 2-micrometers DNA plasmid. ProcNatl Acad Sci USA 79:3428–32.

Jee J, Mizuno T, Kamada K, et al. (2010). Structure and muta-genesis studies of the C-terminal region of licensing factorCdt1 enable the identification of key residues for bindingto replicative helicase Mcm proteins. J Biol Chem285:15931–40.

136 M. W. PARKER ET AL.

Page 33: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Jenkinson ER, Chong JP. (2006). Minichromosome mainten-ance helicase activity is controlled by N- and C-terminalmotifs and requires the ATPase domain helix-2 insert. ProcNatl Acad Sci USA 103:7613–18.

Jiang W, Mcdonald D, Hope TJ, Hunter T. (1999a).Mammalian Cdc7-Dbf4 protein kinase complex is essentialfor initiation of DNA replication. EMBO J 18:5703–13.

Jiang W, Wells NJ, Hunter T. (1999b). Multistep regulation ofDNA replication by Cdk phosphorylation of HsCdc6. ProcNatl Acad Sci USA 96:6193–8.

Kalfalah FM, Berg E, Christensen MO, et al. (2015). Spatio-temporal regulation of the human licensing factor Cdc6 inreplication and mitosis. Cell Cycle 14:1704–15.

Kamimura Y, Masumoto H, Sugino A, Araki H. (1998). Sld2,which interacts with Dpb11 in Saccharomyces cerevisiae, isrequired for chromosomal DNA replication. Mol Cell Biol18:6102–9.

Kamimura Y, Tak YS, Sugino A, Araki H. (2001). Sld3, whichinteracts with Cdc45 (Sld4), functions for chromosomalDNA replication in Saccharomyces cerevisiae. EMBO J20:2097–107.

Kanemaki M, Sanchez-Diaz A, Gambus A, Labib K. (2003).Functional proteomic identification of DNA replication pro-teins by induced proteolysis in vivo. Nature 423:720–4.

Kang S, Warner MD, Bell SP. (2014). Multiple functions forMcm2-7 ATPase motifs during replication initiation. MolCell 55:655–65.

Kang YH, Galal WC, Farina A, et al. (2012). Properties of thehuman Cdc45/Mcm2-7/GINS helicase complex and itsaction with DNA polymerase epsilon in rolling circle DNAsynthesis. Proc Natl Acad Sci USA 109:6042–7.

Kaplan DL. (2000). The 3'-tail of a forked-duplex stericallydetermines whether one or two DNA strands pass throughthe central channel of a replication-fork helicase. J MolBiol 301:285–99.

Kaplan DL, Davey MJ, O'donnell M. (2003). Mcm4,6,7 uses a“pump in ring” mechanism to unwind DNA by steric exclu-sion and actively translocate along a duplex. J Biol Chem278:49171–82.

Kaplan DL, O'donnell M. (2002). DnaB drives DNA branchmigration and dislodges proteins while encircling twoDNA strands. Mol Cell 10:647–57.

Karnani N, Taylor CM, Malhotra A, Dutta A. (2010). Genomicstudy of replication initiation in human chromosomesreveals the influence of transcription regulation and chro-matin structure on origin selection. Mol Biol Cell21:393–404.

Kasiviswanathan R, Shin JH, Melamud E, Kelman Z. (2004).Biochemical characterization of the Methanothermobacterthermautotrophicus minichromosome maintenance (MCM)helicase N-terminal domains. J Biol Chem 279:28358–66.

Kawakami H, Ohashi E, Kanamoto S, et al. (2015). Specificbinding of eukaryotic ORC to DNA replication originsdepends on highly conserved basic residues. Sci Rep5:14929.

Kawasaki Y, Kim HD, Kojima A, et al. (2006). Reconstitution ofSaccharomyces cerevisiae prereplicative complex assemblyin vitro. Genes Cells 11:745–56.

Kelley LA, Mezulis S, Yates CM, et al. (2015). The Phyre2 webportal for protein modeling, prediction and analysis. NatProtoc 10:845–58.

Kelly TJ, Martin GS, Forsburg SL, et al. (1993). The fissionyeast cdc18þ gene product couples S phase to START andmitosis. Cell 74:371–82.

Kelman Z, Lee JK, Hurwitz J. (1999). The single minichromo-some maintenance protein of Methanobacterium ther-moautotrophicum DeltaH contains DNA helicase activity.Proc Natl Acad Sci USA 96:14783–8.

Kerns SL, Torke SJ, Benjamin JM, Mcgarry TJ. (2007). Gemininprevents rereplication during xenopus development. J BiolChem 282:5514–21.

Khayrutdinov BI, Bae WJ, Yun YM, et al. (2009). Structure ofthe Cdt1 C-terminal domain: conservation of the wingedhelix fold in replication licensing factors. Protein Sci18:2252–64.

Kim SM, Zhang DY, Huberman JA. (2001). Multiple redundantsequence elements within the fission yeast ura4 replicationorigin enhancer. BMC Mol Biol 2:1.

Kimura H, Takizawa N, Nozaki N, Sugimoto K. (1995).Molecular cloning of cDNA encoding mouse Cdc21 andCDC46 homologs and characterization of the products:physical interaction between P1(MCM3) and CDC46 pro-teins. Nucleic Acids Res 23:2097–104.

Kitada K, Johnston LH, Sugino T, Sugino A. (1992).Temperature-sensitive cdc7 mutations of Saccharomycescerevisiae are suppressed by the DBF4 gene, which isrequired for the G1/S cell cycle transition. Genetics131:21–9.

Klemm RD, Austin RJ, Bell SP. (1997). Coordinate binding ofATP and origin DNA regulates the ATPase activity of theorigin recognition complex. Cell 88:493–502.

Knapp D, Bhoite L, Stillman DJ, Nasmyth K. (1996). The tran-scription factor Swi5 regulates expression of the cyclin kin-ase inhibitor p40SIC1. Mol Cell Biol 16:5701–7.

Kneissl M, Putter V, Szalay AA, Grummt F. (2003). Interactionand assembly of murine pre-replicative complex proteinsin yeast and mouse cells. J Mol Biol 327:111–28.

Knott SR, Viggiani CJ, Tavare S, Aparicio OM. (2009).Genome-wide replication profiles indicate an expansiverole for Rpd3L in regulating replication initiation timing orefficiency, and reveal genomic loci of Rpd3 function inSaccharomyces cerevisiae. Genes Dev 23:1077–90.

Kondo T, Kobayashi M, Tanaka J, et al. (2004). Rapid degrad-ation of Cdt1 upon UV-induced DNA damage is mediatedby SCFSkp2 complex. J Biol Chem 279:27315–19.

Kong D, Coleman TR, Depamphilis ML. (2003). Xenopus originrecognition complex (ORC) initiates DNA replication prefer-entially at sequences targeted by Schizosaccharomycespombe ORC. EMBO J 22:3441–50.

Kong D, Depamphilis ML. (2001). Site-specific DNA binding ofthe Schizosaccharomyces pombe origin recognition com-plex is determined by the Orc4 subunit. Mol Cell Biol21:8095–103.

Koonin EV. (1993). A common set of conserved motifs in avast variety of putative nucleic acid-dependent ATPasesincluding MCM proteins involved in the initiation ofeukaryotic DNA replication. Nucleic Acids Res 21:2541–7.

Krastanova I, Sannino V, Amenitsch H, et al. (2012). Structuraland functional insights into the DNA replication factorCdc45 reveal an evolutionary relationship to the DHH fam-ily of phosphoesterases. J Biol Chem 287:4121–8.

Kreitz S, Ritzi M, Baack M, Knippers R. (2001). The human ori-gin recognition complex protein 1 dissociates from

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 137

Page 34: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

chromatin during S phase in HeLa cells. J Biol Chem276:6337–42.

Krude T, Musahl C, Laskey RA, Knippers R. (1996). Human rep-lication proteins hCdc21, hCdc46 and P1Mcm3 bind chro-matin uniformly before S-phase and are displaced locallyduring DNA replication. J Cell Sci 109:309–18.

Kubota Y, Mimura S, Nishimoto S, et al. (1995). Identificationof the yeast MCM3-related protein as a component ofXenopus DNA replication licensing factor. Cell 81:601–9.

Kubota Y, Takase Y, Komori Y, et al. (2003). A novel ring-likecomplex of Xenopus proteins essential for the initiation ofDNA replication. Genes Dev 17:1141–52.

Kumagai A, Shevchenko A, Shevchenko A, Dunphy WG.(2010). Treslin collaborates with TopBP1 in triggering theinitiation of DNA replication. Cell 140:349–59.

Kuo AJ, Song J, Cheung P, et al. (2012). The BAH domain ofORC1 links H4K20me2 to DNA replication licensing andMeier-Gorlin syndrome. Nature 484:115–19.

Labib K, Diffley JF, Kearsey SE. (1999). G1-phase and B-typecyclins exclude the DNA-replication factor Mcm4 from thenucleus. Nat Cell Biol 1:415–22.

Labib K, Tercero JA, Diffley JF. (2000). Uninterrupted MCM2-7function required for DNA replication fork progression.Science 288:1643–7.

Lam SK, Ma X, Sing TL, et al. (2013). The PS1 hairpin ofMcm3 is essential for viability and for DNA unwinding invitro. PLoS One 8:e82177.

Landis G, Kelley R, Spradling AC, Tower J. (1997). The k43gene, required for chorion gene amplification and diploidcell chromosome replication, encodes the Drosophilahomolog of yeast origin recognition complex subunit 2.Proc Natl Acad Sci USA 94:3888–92.

Leatherwood J, Lopez-Girona A, Russell P. (1996). Interactionof Cdc2 and Cdc18 with a fission yeast ORC2-like protein.Nature 379:360–3.

Leatherwood J, Vas A. (2003). Connecting ORC and hetero-chromatin: why? Cell Cycle 2:573–5.

Lee C, Hong B, Choi JM, et al. (2004). Structural basis forinhibition of the replication licensing factor Cdt1 by gemi-nin. Nature 430:913–17.

Lee DG, Bell SP. (1997). Architecture of the yeast origin rec-ognition complex bound to origins of DNA replication.Mol Cell Biol 17:7159–68.

Lee DG, Makhov AM, Klemm RD, et al. (2000). Regulation oforigin recognition complex conformation and ATPaseactivity: differential effects of single-stranded and double-stranded DNA binding. EMBO J 19:4774–82.

Lee JK, Hurwitz J. (2000). Isolation and characterization ofvarious complexes of the minichromosome maintenanceproteins of Schizosaccharomyces pombe. J Biol Chem275:18871–8.

Lei M, Kawasaki Y, Young MR, et al. (1997). Mcm2 is a targetof regulation by Cdc7-Dbf4 during the initiation of DNAsynthesis. Genes Dev 11:3365–74.

Leipe DD, Aravind L, Koonin EV. (1999). Did DNA replicationevolve twice independently?. Nucleic Acids Res27:3389–401.

Lenzen CU, Steinmann D, Whiteheart SW, Weis WI. (1998).Crystal structure of the hexamerization domain of N-ethyl-maleimide-sensitive fusion protein. Cell 94:525–36.

Leon RP, Tecklenburg M, Sclafani RA. (2008). Functional con-servation of beta-hairpin DNA binding domains in the

Mcm protein of Methanobacterium thermoautotrophicumand the Mcm5 protein of Saccharomyces cerevisiae.Genetics 179:1757–68.

Leonard AC, Mechali M. (2013). DNA replication origins. ColdSpring Harb Perspect Biol 5:a010116.

Li CJ, Depamphilis ML. (2002). Mammalian Orc1 protein isselectively released from chromatin and ubiquitinated dur-ing the S-to-M transition in the cell division cycle. Mol CellBiol 22:105–16.

Li CJ, Vassilev A, Depamphilis ML. (2004). Role for Cdk1(Cdc2)/cyclin A in preventing the mammalian origin recog-nition complex's largest subunit (Orc1) from binding tochromatin during mitosis. Mol Cell Biol 24:5875–86.

Li JJ, Herskowitz I. (1993). Isolation of ORC6, a component ofthe yeast origin recognition complex by a one-hybrid sys-tem. Science 262:1870–4.

Li N, Zhai Y, Zhang Y, et al. (2015). Structure of the eukary-otic MCM complex at 3.8 Å. Nature 524:186–91.

Li X, Zhao Q, Liao R, et al. (2003). The SCF(Skp2) ubiquitin lig-ase complex interacts with the human replication licensingfactor Cdt1 and regulates Cdt1 degradation. J Biol Chem278:30854–8.

Liang C, Stillman B. (1997). Persistent initiation of DNA repli-cation and chromatin-bound MCM proteins during the cellcycle in cdc6 mutants. Genes Dev 11:3375–86.

Liang C, Weinreich M, Stillman B. (1995). ORC and Cdc6pinteract and determine the frequency of initiation of DNAreplication in the genome. Cell 81:667–76.

Lidonnici MR, Rossi R, Paixao S, et al. (2004). Subnucleardistribution of the largest subunit of the human originrecognition complex during the cell cycle. J Cell Sci117:5221–31.

Liku ME, Nguyen VQ, Rosales AW, et al. (2005). CDK phos-phorylation of a novel NLS-NES module distributedbetween two subunits of the Mcm2-7 complex preventschromosomal rereplication. Mol Biol Cell 16:5026–39.

Lipford JR, Bell SP. (2001). Nucleosomes positioned by ORCfacilitate the initiation of DNA replication. Mol Cell7:21–30.

Lisziewicz J, Godany A, Agoston DV, Kuntzel H. (1988).Cloning and characterization of the Saccharomyces cerevi-siae CDC6 gene. Nucleic Acids Res 16:11507–20.

Liu C, Wu R, Zhou B, et al. (2012a). Structural insights intothe Cdt1-mediated MCM2-7 chromatin loading. NucleicAcids Res 40:3208–17.

Liu E, Li X, Yan F, et al. (2004). Cyclin-dependent kinasesphosphorylate human Cdt1 and induce its degradation.J Biol Chem 279:17283–8.

Liu J, Mcconnell K, Dixon M, Calvi BR. (2012b). Analysis ofmodel replication origins in Drosophila reveals newaspects of the chromatin landscape and its relationship toorigin activity and the prereplicative complex. Mol BiolCell 23:200–12.

Liu J, Smith CL, Deryckere D, et al. (2000). Structure and func-tion of Cdc6/Cdc18: implications for origin recognition andcheckpoint control. Mol Cell 6:637–48.

Liu S, Balasov M, Wang H, et al. (2011). Structural analysis ofhuman Orc6 protein reveals a homology with transcriptionfactor TFIIB. Proc Natl Acad Sci USA 108:7373–8.

Liu W, Pucci B, Rossi M, et al. (2008). Structural analysis ofthe Sulfolobus solfataricus MCM protein N-terminaldomain. Nucleic Acids Res 36:3235–43.

138 M. W. PARKER ET AL.

Page 35: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Loo S, Fox CA, Rine J, et al. (1995). The origin recognitioncomplex in silencing, cell cycle progression, and DNA repli-cation. Mol Biol Cell 6:741–56.

Lubelsky Y, Sasaki T, Kuipers MA, et al. (2011). Pre-replicationcomplex proteins assemble at regions of low nucleosomeoccupancy within the Chinese hamster dihydrofolatereductase initiation zone. Nucleic Acids Res 39:3141–55.

Lutzmann M, Maiorano D, Mechali M. (2006). A Cdt1-geminincomplex licenses chromatin for DNA replication and pre-vents rereplication during S phase in Xenopus. EMBO J25:5764–74.

Lyubimov AY, Costa A, Bleichert F, et al. (2012). ATP-depend-ent conformational dynamics underlie the functional asym-metry of the replicative helicase from a minimalisteukaryote. Proc Natl Acad Sci USA 109:11999–2004.

Macalpine DM, Bell SP. (2005). A genomic view of eukaryoticDNA replication. Chromosome Res 13:309–26.

Macalpine HK, Gordan R, Powell SK, et al. (2010). DrosophilaORC localizes to open chromatin and marks sites of cohe-sin complex loading. Genome Res 20:201–11.

Madine MA, Khoo CY, Mills AD, Laskey RA. (1995). MCM3complex required for cell cycle regulation of DNA replica-tion in vertebrate cells. Nature 375:421–4.

Maine GT, Sinha P, Tye BK. (1984). Mutants of S. cerevisiaedefective in the maintenance of minichromosomes.Genetics 106:365–85.

Maiorano D, Moreau J, Mechali M. (2000). XCDT1 is requiredfor the assembly of pre-replicative complexes in Xenopuslaevis. Nature 404:622–5.

Maiorano D, Rul W, Mechali M. (2004). Cell cycle regulationof the licensing activity of Cdt1 in Xenopus laevis. Exp CellRes 295:138–49.

Makarova KS, Koonin EV, Kelman Z. (2012). The CMG (CDC45/RecJ, MCM, GINS) complex is a conserved component ofthe DNA replication system in all archaea and eukaryotes.Biol Direct 7:7.

Makise M, Takenaka H, Kuwae W, et al. (2003). Kinetics ofATP binding to the origin recognition complex ofSaccharomyces cerevisiae. J Biol Chem 278:46440–5.

Marahrens Y, Stillman B. (1992). A yeast chromosomal originof DNA replication defined by multiple functional ele-ments. Science 255:817–23.

Masai H, Matsumoto S, You Z, et al. (2010). Eukaryoticchromosome DNA replication: where, when, and how?Annu Rev Biochem 79:89–130.

Masai H, Taniyama C, Ogino K, et al. (2006). Phosphorylationof MCM4 by Cdc7 kinase facilitates its interaction withCdc45 on the chromatin. J Biol Chem 281:39249–61.

Masuda T, Mimura S, Takisawa H. (2003). CDK- and Cdc45-dependent priming of the MCM complex on chromatinduring S-phase in Xenopus egg extracts: possible activa-tion of MCM helicase by association with Cdc45. GenesCells 8:145–61.

Masumoto H, Muramatsu S, Kamimura Y, Araki H. (2002). S-Cdk-dependent phosphorylation of Sld2 essential forchromosomal DNA replication in budding yeast. Nature415:651–5.

Matsuno K, Kumano M, Kubota Y, et al. (2006). The N-ter-minal noncatalytic region of Xenopus RecQ4 is requiredfor chromatin binding of DNA polymerase alpha in the ini-tiation of DNA replication. Mol Cell Biol 26:4843–52.

Mcgarry TJ, Kirschner MW. (1998). Geminin, an inhibitor ofDNA replication, is degraded during mitosis. Cell93:1043–53.

Mcgeoch AT, Trakselis MA, Laskey RA, Bell SD. (2005).Organization of the archaeal MCM complex on DNA andimplications for the helicase mechanism. Nat Struct MolBiol 12:756–62.

Mechali M, Kearsey S. (1984). Lack of specific sequencerequirement for DNA replication in Xenopus eggs com-pared with high sequence specificity in yeast. Cell38:55–64.

Mendenhall MD. (1993). An inhibitor of p34CDC28 proteinkinase activity from saccharomyces cerevisiae. Science259:216–19.

Mendez J, Zou-Yang XH, Kim SY, et al. (2002). Human originrecognition complex large subunit is degraded by ubiqui-tin-mediated proteolysis after initiation of DNA replication.Mol Cell 9:481–91.

Micklem G, Rowley A, Harwood J, et al. (1993). Yeast originrecognition complex is involved in DNA replication andtranscriptional silencing. Nature 366:87–9.

Mihaylov IS, Kondo T, Jones L, et al. (2002). Control of DNAreplication and chromosome ploidy by geminin and cyclinA. Mol Cell Biol 22:1868–80.

Miller JM, Arachea BT, Epling LB, Enemark EJ. (2014). Analysisof the crystal structure of an active MCM hexamer. Elife3:e03433.

Mimura S, Seki T, Tanaka S, Diffley JF. (2004).Phosphorylation-dependent binding of mitotic cyclins toCdc6 contributes to DNA replication control. Nature431:1118–23.

Miotto B, Struhl K. (2010). HBO1 histone acetylase activity isessential for DNA replication licensing and inhibited byGeminin. Mol Cell 37:57–66.

Mizushima T, Takahashi N, Stillman B. (2000). Cdc6p modu-lates the structure and DNA binding activity of the originrecognition complex in vitro. Genes Dev 14:1631–41.

Moir D, Stewart SE, Osmond BC, Botstein D. (1982). Cold-sen-sitive cell-division-cycle mutants of yeast: isolation, proper-ties, and pseudoreversion studies. Genetics 100:547–63.

Moll T, Tebb G, Surana U, et al. (1991). The role of phosphor-ylation and the CDC28 protein kinase in cell cycle-regu-lated nuclear import of the S. cerevisiae transcription factorSWI5. Cell 66:743–58.

Montagnoli A, Valsasina B, Brotherton D, et al. (2006).Identification of Mcm2 phosphorylation sites by S-phase-regulating kinases. J Biol Chem 281:10281–90.

Moon KY, Kong D, Lee JK, et al. (1999). Identification andreconstitution of the origin recognition complex fromSchizosaccharomyces pombe. Proc Natl Acad Sci USA96:12367–72.

Moreau MJ, Mcgeoch AT, Lowe AR, et al. (2007). ATPase sitearchitecture and helicase mechanism of an archaeal MCM.Mol Cell 28:304–14.

Moyer SE, Lewis PW, Botchan MR. (2006). Isolation of theCdc45/Mcm2-7/GINS (CMG) complex, a candidate for theeukaryotic DNA replication fork helicase. Proc Natl AcadSci USA 103:10236–41.

Muller P, Park S, Shor E, et al. (2010). The conserved bromo-adjacent homology domain of yeast Orc1 functions in theselection of DNA replication origins within chromatin.Genes Dev 24:1418–33.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 139

Page 36: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Muramatsu S, Hirai K, Tak YS, et al. (2010). CDK-dependentcomplex formation between replication proteins Dpb11,Sld2, Pol (epsilon}, and GINS in budding yeast. Genes Dev24:602–12.

Musahl C, Schulte D, Burkhart R, Knippers R. (1995). A humanhomologue of the yeast replication protein Cdc21.Interactions with other Mcm proteins. Eur J Biochem230:1096–101.

Muzi-Falconi M, Kelly TJ. (1995). Orp1, a member of theCdc18/Cdc6 family of S-phase regulators, is homologousto a component of the origin recognition complex. ProcNatl Acad Sci USA 92:12475–9.

Narasimhachar Y, Coue M. (2009). Geminin stabilizes Cdt1during meiosis in Xenopus oocytes. J Biol Chem284:27235–42.

Natale DA, Li CJ, Sun WH, Depamphilis ML. (2000). Selectiveinstability of Orc1 protein accounts for the absence offunctional origin recognition complexes during the M-G(1)transition in mammals. EMBO J 19:2728–38.

Neuwald AF, Aravind L, Spouge JL, Koonin EV. (1999). AAAþ:a class of chaperone-like ATPases associated with theassembly, operation, and disassembly of protein com-plexes. Genome Res 9:27–43.

Nguyen VQ, Co C, Li JJ. (2001). Cyclin-dependent kinases pre-vent DNA re-replication through multiple mechanisms.Nature 411:1068–73.

Nieduszynski CA, Knox Y, Donaldson AD. (2006). Genome-wide identification of replication origins in yeast by com-parative genomics. Genes Dev 20:1874–9.

Nishitani H, Lygerou Z, Nishimoto T. (2004). Proteolysis ofDNA replication licensing factor Cdt1 in S-phase is per-formed independently of geminin through its N-terminalregion. J Biol Chem 279:30807–16.

Nishitani H, Lygerou Z, Nishimoto T, Nurse P. (2000). TheCdt1 protein is required to license DNA for replication infission yeast. Nature 404:625–8.

Nishitani H, Sugimoto N, Roukos V, et al. (2006). Two E3 ubi-quitin ligases, SCF-Skp2 and DDB1-Cul4, target humanCdt1 for proteolysis. EMBO J 25:1126–36.

Nishitani H, Taraviras S, Lygerou Z, Nishimoto T. (2001). Thehuman licensing factor for DNA replication Cdt1 accumu-lates in G1 and is destabilized after initiation of S-phase. JBiol Chem 276:44905–11.

Noguchi K, Vassilev A, Ghosh S, et al. (2006). The BAHdomain facilitates the ability of human Orc1 protein toactivate replication origins in vivo. EMBO J 25:5372–82.

Nougarede R, Della Seta F, Zarzov P, Schwob E. (2000).Hierarchy of S-phase-promoting factors: yeast Dbf4-Cdc7kinase requires prior S-phase cyclin-dependent kinase acti-vation. Mol Cell Biol 20:3795–806.

Ohlenschlager O, Kuhnert A, Schneider A, et al. (2012). TheN-terminus of the human RecQL4 helicase is a homeodo-main-like DNA interaction motif. Nucleic Acids Res40:8309–24.

Ohta S, Tatsumi Y, Fujita M, et al. (2003). The ORC1 cycle inhuman cells: II. Dynamic changes in the human ORC com-plex during the cell cycle. J Biol Chem 278:41535–40.

Okuno Y, Satoh H, Sekiguchi M, Masukata H. (1999).Clustered adenine/thymine stretches are essential for func-tion of a fission yeast replication origin. Mol Cell Biol19:6699–709.

Oshiro G, Owens JC, Shellman Y, et al. (1999). Cell cycle con-trol of Cdc7p kinase activity through regulation of Dbf4pstability. Mol Cell Biol 19:4888–96.

Pacek M, Tutter AV, Kubota Y, et al. (2006). Localization ofMCM2-7, Cdc45, and GINS to the site of DNA unwindingduring eukaryotic DNA replication. Mol Cell 21:581–7.

Pacek M, Walter JC. (2004). A requirement for MCM7 andCdc45 in chromosome unwinding during eukaryotic DNAreplication. EMBO J 23:3667–76.

Pak DT, Pflumm M, Chesnokov I, et al. (1997). Association ofthe origin recognition complex with heterochromatin andHP1 in higher eukaryotes. Cell 91:311–23.

Palacios Debeer MA, Muller U, Fox CA. (2003). DifferentialDNA affinity specifies roles for the origin recognition com-plex in budding yeast heterochromatin. Genes Dev17:1817–22.

Palmer RE, Hogan E, Koshland D. (1990). Mitotic transmissionof artificial chromosomes in cdc mutants of the yeast,Saccharomyces cerevisiae. Genetics 125:763–74.

Perez-Arnaiz P, Kaplan DL. (2016). An Mcm10 mutant defect-ive in ssDNA binding shows defects in dna replication initi-ation. J Mol Biol 428:4608–25.

Perkins G, Diffley JF. (1998). Nucleotide-dependent prereplica-tive complex assembly by Cdc6p, a homolog of eukaryoticand prokaryotic clamp-loaders. Mol Cell 2:23–32.

Perkins G, Drury LS, Diffley JF. (2001). Separate SCF(CDC4)recognition elements target Cdc6 for proteolysis in Sphase and mitosis. EMBO J 20:4836–45.

Petersen BO, Lukas J, Sorensen CS, et al. (1999).Phosphorylation of mammalian CDC6 by cyclinA/CDK2 regulates its subcellular localization. EMBO J18:396–410.

Petojevic T, Pesavento JJ, Costa A, et al. (2015). Cdc45 (celldivision cycle protein 45) guards the gate of the eukaryotereplisome helicase stabilizing leading strand engagement.Proc Natl Acad Sci USA 112:E249–58.

Piatti S, Lengauer C, Nasmyth K. (1995). Cdc6 is an unstableprotein whose de novo synthesis in G1 is important forthe onset of S phase and for preventing a 'reductional'anaphase in the budding yeast Saccharomyces cerevisiae.EMBO J 14:3788–99.

Poplawski A, Grabowski B, Long SE, Kelman Z. (2001). Thezinc finger domain of the archaeal minichromosome main-tenance protein is required for helicase activity. J BiolChem 276:49371–7.

Powell SK, Macalpine HK, Prinz JA, et al. (2015). Dynamicloading and redistribution of the Mcm2-7 helicase complexthrough the cell cycle. EMBO J 34:531–43.

Prasanth SG, Prasanth KV, Siddiqui K, et al. (2004). HumanOrc2 localizes to centrosomes, centromeres and hetero-chromatin during chromosome inheritance. EMBO J23:2651–63.

Prasanth SG, Shen Z, Prasanth KV, Stillman B. (2010). Humanorigin recognition complex is essential for HP1 binding tochromatin and heterochromatin organization. Proc NatlAcad Sci USA 107:15093–8.

Prioleau MN. (2009). CpG islands: starting blocks for replica-tion and transcription. PLoS Genet 5:e1000454.

Prokhorova TA, Blow JJ. (2000). Sequential MCM/P1 subcom-plex assembly is required to form a heterohexamer withreplication licensing activity. J Biol Chem 275:2491–8.

140 M. W. PARKER ET AL.

Page 37: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Pucci B, De Felice M, Rocco M, et al. (2007). Modular organ-ization of the Sulfolobus solfataricus mini-chromosomemaintenance protein. J Biol Chem 282:12574–82.

Pucci B, De Felice M, Rossi M, et al. (2004). Amino acids ofthe Sulfolobus solfataricus mini-chromosome mainten-ance-like DNA helicase involved in DNA binding/remodel-ing. J Biol Chem 279:49222–8.

Putnam CD, Clancy SB, Tsuruta H, et al. (2001). Structure andmechanism of the RuvB Holliday junction branch migra-tion motor. J Mol Biol 311:297–310.

Quinn LM, Herr A, Mcgarry TJ, Richardson H. (2001). TheDrosophila Geminin homolog: roles for Geminin in limitingDNA replication, in anaphase and in neurogenesis. GenesDev 15:2741–54.

Quintana DG, Hou Z, Thome KC, et al. (1997). Identification ofHsORC4, a member of the human origin of replication rec-ognition complex. J Biol Chem 272:28247–51.

Ramer MD, Suman ES, Richter H, et al. (2013). Dbf4 and Cdc7proteins promote DNA replication through interactionswith distinct Mcm2-7 protein subunits. J Biol Chem288:14926–35.

Ramey CJ, Sclafani RA. (2014). Functional conservation of thepre-sensor one beta-finger hairpin (PS1-hp) structures inmini-chromosome maintenance proteins of Saccharomycescerevisiae and archaea. G3 (Bethesda) 4:1319–26.

Randell JC, Bowers JL, Rodriguez HK, Bell SP. (2006).Sequential ATP hydrolysis by Cdc6 and ORC directs load-ing of the Mcm2-7 helicase. Mol Cell 21:29–39.

Randell JC, Fan A, Chan C, et al. (2010). Mec1 is one of mul-tiple kinases that prime the Mcm2-7 helicase for phos-phorylation by Cdc7. Mol Cell 40:353–63.

Ranjan A, Gossen M. (2006). A structural role for ATP in theformation and stability of the human origin recognitioncomplex. Proc Natl Acad Sci USA 103:4864–9.

Rao H, Marahrens Y, Stillman B. (1994). Functional conserva-tion of multiple elements in yeast chromosomal replica-tors. Mol Cell Biol 14:7643–51.

Rao H, Stillman B. (1995). The origin recognition complexinteracts with a bipartite DNA binding site within yeastreplicators. Proc Natl Acad Sci USA 92:2224–8.

Remus D, Beall EL, Botchan MR. (2004). DNA topology, notDNA sequence, is a critical determinant for DrosophilaORC-DNA binding. EMBO J 23:897–907.

Remus D, Beuron F, Tolun G, et al. (2009). Concerted loadingof Mcm2-7 double hexamers around DNA during DNA rep-lication origin licensing. Cell 139:719–30.

Rice JC, Nishioka K, Sarma K, et al. (2002). Mitotic-specificmethylation of histone H4 Lys 20 follows increased PR-Set7 expression and its localization to mitotic chromo-somes. Genes Dev 16:2225–30.

Ricke RM, Bielinsky AK. (2004). Mcm10 regulates the stabilityand chromatin association of DNA polymerase-alpha. MolCell 16:173–85.

Ritzi M, Baack M, Musahl C, et al. (1998). Human minichromo-some maintenance proteins and human origin recognitioncomplex 2 protein oil chromatin. J Biol Chem 273:24543–9.

Robinson NP, Bell SD. (2007). Extrachromosomal element cap-ture and the evolution of multiple replication origins inarchaeal chromosomes. Proc Natl Acad Sci USA104:5806–11.

Romanowski P, Madine MA, Laskey RA. (1996a). XMCM7, anovel member of the Xenopus MCM family, interacts with

XMCM3 and colocalizes with it throughout replication.Proc Natl Acad Sci USA 93:10189–94.

Romanowski P, Madine MA, Rowles A, et al. (1996b). TheXenopus origin recognition complex is essential for DNAreplication and MCM binding to chromatin. Curr Biol6:1416–25.

Rothenberg E, Trakselis MA, Bell SD, Ha T. (2007). MCMforked substrate specificity involves dynamic interactionwith the 5'-tail. J Biol Chem 282:34229–34.

Rowles A, Chong JP, Brown L, et al. (1996). Interactionbetween the origin recognition complex and the replica-tion licensing system in Xenopus. Cell 87:287–96.

Rowles A, Tada S, Blow JJ. (1999). Changes in association ofthe Xenopus origin recognition complex with chromatinon licensing of replication origins. J Cell Sci 112:2011–18.

Royzman I, Austin RJ, Bosco G, et al. (1999). ORC localizationin Drosophila follicle cells and the effects of mutations indE2F and dDP. Genes Dev 13:827–40.

Saha P, Chen J, Thome KC, et al. (1998). Human CDC6/Cdc18associates with Orc1 and cyclin-cdk and is selectively elimi-nated from the nucleus at the onset of S phase. Mol CellBiol 18:2758–67.

Sakakibara N, Kasiviswanathan R, Melamud E, et al. (2008).Coupling of DNA binding and helicase activity is mediatedby a conserved loop in the MCM protein. Nucleic AcidsRes 36:1309–20.

Samel SA, Fernandez-Cid A, Sun J, et al. (2014). A uniqueDNA entry gate serves for regulated loading of the eukary-otic replicative helicase MCM2-7 onto DNA. Genes Dev28:1653–66.

Samson RY, Abeyrathne PD, Bell SD. (2016). Mechanism ofarchaeal mcm helicase recruitment to dna replication ori-gins. Mol Cell 61:287–96.

Samson RY, Xu Y, Gadelha C, et al. (2013). Specificity andfunction of archaeal DNA replication initiator proteins. CellRep 3:485–96.

Sanchez-Pulido L, Diffley JF, Ponting CP. (2010). Homologyexplains the functional similarities of Treslin/Ticrr and Sld3.Curr Biol 20:R509–10.

Sanchez-Pulido L, Ponting CP. (2011). Cdc45: the missingRecJ ortholog in eukaryotes?. Bioinformatics 27:1885–8.

Sangrithi MN, Bernal JA, Madine M, et al. (2005). Initiation ofDNA replication requires the RECQL4 protein mutated inRothmund-Thomson syndrome. Cell 121:887–98.

Sansam CL, Cruz NM, Danielian PS, et al. (2010). A vertebrategene, ticrr, is an essential checkpoint and replication regu-lator. Genes Dev 24:183–94.

Santocanale C, Diffley JF. (1996). ORC- and Cdc6-dependentcomplexes at active and inactive chromosomal replicationorigins in Saccharomyces cerevisiae. EMBO J 15:6671–9.

Saxena S, Yuan P, Dhar SK, et al. (2004). A dimerized coiled-coil domain and an adjoining part of geminin interactwith two sites on Cdt1 for replication inhibition. Mol Cell15:245–58.

Schepers A, Diffley JF. (2001). Mutational analysis of con-served sequence motifs in the budding yeast Cdc6 protein.J Mol Biol 308:597–608.

Schneider BL, Yang QH, Futcher AB. (1996). Linkage of repli-cation to start by the Cdk inhibitor Sic1. Science272:560–2.

Schwacha A, Bell SP. (2001). Interactions between two cata-lytically distinct MCM subgroups are essential for

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 141

Page 38: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

coordinated ATP hydrolysis and DNA replication. Mol Cell8:1093–104.

Schwob E, Bohm T, Mendenhall MD, Nasmyth K. (1994). TheB-type cyclin kinase inhibitor p40SIC1 controls the G1 to Stransition in S. cerevisiae. Cell 79:233–44.

Segurado M, De Luis A, Antequera F. (2003). Genome-widedistribution of DNA replication origins at Aþ T-rich islandsin Schizosaccharomyces pombe. EMBO Rep 4:1048–53.

Seki T, Diffley JF. (2000). Stepwise assembly of initiation pro-teins at budding yeast replication origins in vitro. ProcNatl Acad Sci USA 97:14115–20.

Semple JW, Da-Silva LF, Jervis EJ, et al. (2006). An essentialrole for Orc6 in DNA replication through maintenance ofpre-replicative complexes. EMBO J 25:5150–8.

Senga T, Sivaprasad U, Zhu W, et al. (2006). PCNA is a cofac-tor for Cdt1 degradation by CUL4/DDB1-mediated N-ter-minal ubiquitination. J Biol Chem 281:6246–52.

Sequeira-Mendes J, Diaz-Uriarte R, Apedaile A, et al.(2009). Transcription initiation activity sets replicationorigin efficiency in mammalian cells. PLoS Genet5:e1000446.

Shechter D, Ying CY, Gautier J. (2004). DNA unwinding is anMcm complex-dependent and ATP hydrolysis-dependentprocess. J Biol Chem 279:45586–93.

Shechter DF, Ying CY, Gautier J. (2000). The intrinsic DNAhelicase activity of methanobacterium thermoautotrophi-cum delta H minichromosome maintenance protein. J BiolChem 275:15049–59.

Shen Z, Chakraborty A, Jain A, et al. (2012). Dynamic associ-ation of ORCA with prereplicative complex componentsregulates DNA replication initiation. Mol Cell Biol32:3107–20.

Shen Z, Sathyan KM, Geng Y, et al. (2010). A WD-repeat pro-tein stabilizes ORC binding to chromatin. Mol Cell40:99–111.

Sheu YJ, Stillman B. (2006). Cdc7-Dbf4 phosphorylates MCMproteins via a docking site-mediated mechanism to pro-mote S phase progression. Mol Cell 24:101–13.

Sheu YJ, Stillman B. (2010). The Dbf4-Cdc7 kinase promotes Sphase by alleviating an inhibitory activity in Mcm4. Nature463:113–17.

Shor E, Warren CL, Tietjen J, et al. (2009). The origin recogni-tion complex interacts with a subset of metabolic genestightly linked to origins of replication. PLoS Genet5:e1000755.

Shore D, Stillman DJ, Brand AH, Nasmyth KA. (1987).Identification of silencer binding proteins from yeast: pos-sible roles in SIR control and DNA replication. EMBO J6:461–7.

Siddiqui K, On KF, Diffley JF. (2013). Regulating DNA replica-tion in eukarya. Cold Spring Harb Perspect Biol 5. doi:10.1101/cshperspect.a012930

Siddiqui K, Stillman B. (2007). ATP-dependent assembly ofthe human origin recognition complex. J Biol Chem282:32370–83.

Simon AC, Sannino V, Costanzo V, Pellegrini L. (2016).Structure of human Cdc45 and implications for CMG heli-case function. Nat Commun 7:11638.

Simon AC, Zhou JC, Perera RL, et al. (2014). A Ctf4 trimercouples the CMG helicase to DNA polymerase a in theeukaryotic replisome. Nature 510:293–7.

Simpson RT. (1990). Nucleosome positioning can affect thefunction of a cis-acting DNA element in vivo. Nature343:387–9.

Singleton MR, Morales R, Grainge I, et al. (2004).Conformational changes induced by nucleotide binding inCdc6/ORC from Aeropyrum pernix. J Mol Biol 343:547–57.

Speck C, Chen Z, Li H, Stillman B. (2005). ATPase-dependentcooperative binding of ORC and Cdc6 to origin DNA. NatStruct Mol Biol 12:965–71.

Speck C, Stillman B. (2007). Cdc6 ATPase activity regulatesORC x Cdc6 stability and the selection of specific DNAsequences as origins of DNA replication. J Biol Chem282:11705–14.

Stinchcomb DT, Struhl K, Davis RW. (1979). Isolation andcharacterisation of a yeast chromosomal replicator. Nature282:39–43.

Su TT, Feger G, O'farrell PH. (1996). Drosophila MCM proteincomplexes. Mol Biol Cell 7:319–29.

Sugimoto N, Kitabayashi I, Osano S, et al. (2008).Identification of novel human Cdt1-binding proteins by aproteomics approach: proteolytic regulation by APC/CCdh1. Mol Biol Cell 19:1007–21.

Sugimoto N, Tatsumi Y, Tsurumi T, et al. (2004). Cdt1 phos-phorylation by cyclin A-dependent kinases negatively reg-ulates its function without affecting geminin binding. JBiol Chem 279:19691–7.

Sun J, Evrin C, Samel SA, et al. (2013). Cryo-EM structure of ahelicase loading intermediate containing ORC-Cdc6-Cdt1-MCM2-7 bound to DNA. Nat Struct Mol Biol 20:944–51.

Sun J, Fernandez-Cid A, Riera A, et al. (2014). Structural andmechanistic insights into Mcm2-7 double-hexamer assem-bly and function. Genes Dev 28:2291–303.

Sun J, Kawakami H, Zech J, et al. (2012). Cdc6-induced con-formational changes in ORC bound to origin DNA revealedby cryo-electron microscopy. Structure 20:534–44.

Sun WH, Coleman TR, Depamphilis ML. (2002). Cell cycle-dependent regulation of the association between originrecognition proteins and somatic cell chromatin. EMBO J21:1437–46.

Sweder KS, Rhode PR, Campbell JL. (1988). Purification andcharacterization of proteins that bind to yeast ARSs. J BiolChem 263:17270–7.

Szambowska A, Tessmer I, Kursula P, et al. (2014). DNA bind-ing properties of human Cdc45 suggest a function asmolecular wedge for DNA unwinding. Nucleic Acids Res42:2308–19.

Tada S, Chong JP, Mahbubani HM, Blow JJ. (1999). The RLF-Bcomponent of the replication licensing system is distinctfrom Cdc6 and functions after Cdc6 binds to chromatin.Curr Biol 9:211–14.

Tada S, Li A, Maiorano D, et al. (2001). Repression of originassembly in metaphase depends on inhibition of RLF-B/Cdt1 by geminin. Nat Cell Biol 3:107–13.

Takahashi N, Tsutsumi S, Tsuchiya T, et al. (2002). Functionsof sensor 1 and sensor 2 regions of Saccharomyces cerevi-siae Cdc6p in vivo and in vitro. J Biol Chem 277:16033–40.

Takahashi TS, Wigley DB, Walter JC. (2005). Pumps, paradoxesand ploughshares: mechanism of the MCM2-7 DNA heli-case. Trends Biochem Sci 30:437–44.

Takara TJ, Bell SP. (2011). Multiple Cdt1 molecules act ateach origin to load replication-competent Mcm2-7 heli-cases. EMBO J 30:4885–96.

142 M. W. PARKER ET AL.

Page 39: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Takayama Y, Kamimura Y, Okawa M, et al. (2003). GINS, anovel multiprotein complex required for chromosomalDNA replication in budding yeast. Genes Dev 17:1153–65.

Takeda DY, Shibata Y, Parvin JD, Dutta A. (2005). Recruitmentof ORC or CDC6 to DNA is sufficient to create an artificialorigin of replication in mammalian cells. Genes Dev19:2827–36.

Tanaka S, Diffley JF. (2002). Interdependent nuclear accumu-lation of budding yeast Cdt1 and Mcm2-7 during G1phase. Nat Cell Biol 4:198–207.

Tanaka S, Komeda Y, Umemori T, et al. (2013). Efficient initi-ation of DNA replication in eukaryotes requires Dpb11/TopBP1-GINS interaction. Mol Cell Biol 33:2614–22.

Tanaka S, Nakato R, Katou Y, et al. (2011a). Origin associationof Sld3, Sld7, and Cdc45 proteins is a key step for deter-mination of origin-firing timing. Curr Biol 21:2055–63.

Tanaka S, Umemori T, Hirai K, et al. (2007). CDK-dependentphosphorylation of Sld2 and Sld3 initiates DNA replicationin budding yeast. Nature 445:328–32.

Tanaka T, Umemori T, Endo S, et al. (2011b). Sld7, an Sld3-associated protein required for efficient chromosomal DNAreplication in budding yeast. EMBO J 30:2019–30.

Tardat M, Brustel J, Kirsh O, et al. (2010). The histone H4 Lys20 methyltransferase PR-Set7 regulates replication originsin mammalian cells. Nat Cell Biol 12:1086–93.

Tatsumi Y, Ezura K, Yoshida K, et al. (2008). Involvement ofhuman ORC and TRF2 in pre-replication complex assemblyat telomeres. Genes Cells 13:1045–59.

Tatsumi Y, Ohta S, Kimura H, et al. (2003). The ORC1 cycle inhuman cells: I. cell cycle-regulated oscillation of humanORC1. J Biol Chem 278:41528–34.

Teer JK, Dutta A. (2008). Human Cdt1 lacking the evolutionar-ily conserved region that interacts with MCM2-7 is capableof inducing re-replication. J Biol Chem 283:6817–25.

Theis JF, Newlon CS. (1994). Domain B of ARS307 containstwo functional elements and contributes to chromosomalreplication origin function. Mol Cell Biol 14:7652–9.

Thoma F, Bergman LW, Simpson RT. (1984). Nuclease diges-tion of circular TRP1ARS1 chromatin reveals positionednucleosomes separated by nuclease-sensitive regions.J Mol Biol 177:715–33.

Thomae AW, Pich D, Brocher J, et al. (2008). Interactionbetween HMGA1a and the origin recognition complex cre-ates site-specific replication origins. Proc Natl Acad SciUSA 105:1692–7.

Thomer M, May NR, Aggarwal BD, et al. (2004). Drosophiladouble-parked is sufficient to induce re-replication duringdevelopment and is regulated by cyclin E/CDK2.Development 131:4807–18.

Thommes P, Kubota Y, Takisawa H, Blow JJ. (1997). The RLF-M component of the replication licensing system formscomplexes containing all six MCM/P1 polypeptides. EMBOJ 16:3312–19.

Ticau S, Friedman LJ, Ivica NA, et al. (2015). Single-moleculestudies of origin licensing reveal mechanisms ensuringbidirectional helicase loading. Cell 161:513–25.

Todorov IT, Attaran A, Kearsey SE. (1995). BM28, a humanmember of the MCM2-3-5 family, is displaced from chro-matin during DNA replication. J Cell Biol 129:1433–45.

Treisman JE, Follette PJ, O'farrell PH, Rubin GM. (1995). Cellproliferation and DNA replication defects in a DrosophilaMCM2 mutant. Genes Dev 9:1709–15.

Triolo T, Sternglanz R. (1996). Role of interactions betweenthe origin recognition complex and SIR1 in transcriptionalsilencing. Nature 381:251–3.

Tsakraklides V, Bell SP. (2010). Dynamics of pre-replicativecomplex assembly. J Biol Chem 285:9437–43.

Tsuji T, Ficarro SB, Jiang W. (2006). Essential role of phos-phorylation of MCM2 by Cdc7/Dbf4 in the initiation ofDNA replication in mammalian cells. Mol Biol Cell17:4459–72.

Tsunematsu T, Takihara Y, Ishimaru N, et al. (2013). Aurora-Acontrols pre-replicative complex assembly and DNA repli-cation by stabilizing geminin in mitosis. Nat Commun4:1885.

Tsuyama T, Tada S, Watanabe S, et al. (2005). Licensing forDNA replication requires a strict sequential assembly ofCdc6 and Cdt1 onto chromatin in Xenopus egg extracts.Nucleic Acids Res 33:765–75.

Tugal T, Zou-Yang XH, Gavin K, et al. (1998). The Orc4p andOrc5p subunits of the Xenopus and human origin recogni-tion complex are related to Orc1p and Cdc6p. J Biol Chem273:32421–9.

Valton AL, Hassan-Zadeh V, Lema I, et al. (2014). G4 motifsaffect origin positioning and efficiency in two vertebratereplicators. EMBO J 33:732–46.

Van Deursen F, Sengupta S, De Piccoli G, et al. (2012).Mcm10 associates with the loaded DNA helicase at replica-tion origins and defines a novel step in its activation.EMBO J 31:2195–206.

Vas A, Mok W, Leatherwood J. (2001). Control of DNA rerepli-cation via Cdc2 phosphorylation sites in the origin recog-nition complex. Mol Cell Biol 21:5767–77.

Vashee S, Cvetic C, Lu W, et al. (2003). Sequence-independentDNA binding and replication initiation by the human ori-gin recognition complex. Genes Dev 17:1894–908.

Vashee S, Simancek P, Challberg MD, Kelly TJ. (2001).Assembly of the human origin recognition complex. J BiolChem 276:26666–73.

Villa F, Simon AC, Ortiz Bazan MA, et al. (2016). Ctf4 is a hubin the eukaryotic replisome that links multiple CIP-box pro-teins to the cmg helicase. Mol Cell 63:385–96.

Vogelauer M, Rubbi L, Lucas I, et al. (2002). Histone acetyl-ation regulates the time of replication origin firing. MolCell 10:1223–33.

Waga S, Zembutsu A. (2006). Dynamics of DNA binding ofreplication initiation proteins during de novo formation ofpre-replicative complexes in Xenopus egg extracts. J BiolChem 281:10926–34.

Walter JC. (2000). Evidence for sequential action of cdc7 andcdk2 protein kinases during initiation of DNA replication inXenopus egg extracts. J Biol Chem 275:39773–8.

Wang B, Feng L, Hu Y, et al. (1999). The essential role ofSaccharomyces cerevisiae CDC6 nucleotide-binding site incell growth, DNA synthesis, and Orc1 association. J BiolChem 274:8291–8.

Wei L, Zhao X. (2016). A new MCM modification cycle regu-lates DNA replication initiation. Nat Struct Mol Biol23:209–16.

Wei Z, Liu C, Wu X, et al. (2010). Characterization and struc-ture determination of the Cdt1 binding domain of humanminichromosome maintenance (Mcm) 6. J Biol Chem285:12469–73.

CRITICAL REVIEWS IN BIOCHEMISTRY AND MOLECULAR BIOLOGY 143

Page 40: Electronic Delivery Cover Sheet - CALTECHchemistry.caltech.edu/courses/ch111/R4B.pdf · 2019-01-07 · somes possess multiple origins (Wu et al., 2014b), as do all eukaryotic genomes

Weinreich M, Liang C, Stillman B. (1999). The Cdc6p nucleo-tide-binding motif is required for loading mcm proteinsonto chromatin. Proc Natl Acad Sci USA 96:441–6.

Weinreich M, Stillman B. (1999). Cdc7p-Dbf4p kinase binds tochromatin during S phase and is regulated by both theAPC and the RAD53 checkpoint pathway. EMBO J18:5334–46.

Whittaker AJ, Royzman I, Orr-Weaver TL. (2000). Drosophiladouble parked: a conserved, essential replication proteinthat colocalizes with the origin recognition complex andlinks DNA replication with mitosis and the down-regula-tion of S phase transcripts. Genes Dev 14:1765–76.

Wiedemann C, Szambowska A, Hafner S, et al. (2015).Structure and regulatory role of the C-terminal wingedhelix domain of the archaeal minichromosome mainten-ance complex. Nucleic Acids Res 43:2958–67.

Wilmes GM, Archambault V, Austin RJ, et al. (2004).Interaction of the S-phase cyclin Clb5 with an “RXL” dock-ing sequence in the initiator protein Orc6 provides an ori-gin-localized replication control switch. Genes Dev18:981–91.

Wohlschlegel JA, Dhar SK, Prokhorova TA, et al. (2002).Xenopus Mcm10 binds to origins of DNA replication afterMcm2-7 and stimulates origin binding of Cdc45. Mol Cell9:233–40.

Wohlschlegel JA, Dwyer BT, Dhar SK, et al. (2000). Inhibitionof eukaryotic DNA replication by geminin binding to Cdt1.Science 290:2309–12.

Wong HM, Huppert JL. (2009). Stable G-quadruplexes arefound outside nucleosome-bound regions. Mol Biosyst5:1713–19.

Wu M, Lu W, Santos RE, et al. (2014a). Geminin inhibits a latestep in the formation of human pre-replicative complexes.J Biol Chem 289:30810–21.

Wu R, Wang J, Liang C. (2012). Cdt1p, through its interactionwith Mcm6p, is required for the formation, nuclear accu-mulation and chromatin loading of the MCM complex.J Cell Sci 125:209–19.

Wu Z, Liu J, Yang H, Xiang H. (2014b). DNA replication originsin archaea. Front Microbiol 5:179.

Wuarin J, Buck V, Nurse P, Millar JB. (2002). Stable associationof mitotic cyclin B/Cdc2 to replication origins preventsendoreduplication. Cell 111:419–31.

Wyrick JJ, Aparicio JG, Chen T, et al. (2001). Genome-widedistribution of ORC and MCM proteins in S. cerevisiae:high-resolution mapping of replication origins. Science294:2357–60.

Xu J, Yanagisawa Y, Tsankov AM, et al. (2012). Genome-wideidentification and characterization of replication origins bydeep sequencing. Genome Biol 13:R27.

Xu W, Aparicio JG, Aparicio OM, Tavare S. (2006). Genome-wide mapping of ORC and Mcm2p binding sites on tilingarrays and identification of essential ARS consensussequences in S. cerevisiae. BMC Genomics 7:276.

Xu Y, Gristwood T, Hodgson B, et al. (2016). Archaeal ortho-logs of Cdc45 and GINS form a stable complex that stimu-lates the helicase activity of MCM. Proc Natl Acad Sci USA113:13390–5.

Yabuuchi H, Yamada Y, Uchida T, et al. (2006). Orderedassembly of Sld3, GINS and Cdc45 is distinctly regulated

by DDK and CDK for activation of replication origins.EMBO J 25:4663–74.

Yan H, Gibson S, Tye BK. (1991). Mcm2 and Mcm3, two pro-teins important for ARS activity, are related in structureand function. Genes Dev 5:944–57.

Yanagi K, Mizuno T, You Z, Hanaoka F. (2002). Mouse gemi-nin inhibits not only Cdt1-MCM6 interactions but also anovel intrinsic Cdt1 DNA binding activity. J Biol Chem277:40871–80.

Yardimci H, Loveland AB, Habuchi S, et al. (2010). Uncouplingof sister replisomes during eukaryotic DNA replication. MolCell 40:834–40.

Yeeles JT, Deegan TD, Janska A, et al. (2015). Regulatedeukaryotic DNA replication origin firing with purified pro-teins. Nature 519:431–5.

Ying CY, Gautier J. (2005). The ATPase activity of MCM2-7 isdispensable for pre-RC assembly but is required for DNAunwinding. EMBO J 24:4334–44.

Yoshida K, Takisawa H, Kubota Y. (2005). Intrinsic nuclearimport activity of geminin is essential to prevent re-initi-ation of DNA replication in Xenopus eggs. Genes Cells10:63–73.

You Z, Ishimi Y, Mizuno T, et al. (2003). Thymine-rich single-stranded DNA activates Mcm4/6/7 helicase on Y-fork andbubble-like substrates. EMBO J 22:6148–60.

You Z, Masai H. (2008). Cdt1 forms a complex with the mini-chromosome maintenance protein (MCM) and activates itshelicase activity. J Biol Chem 283:24469–77.

Yuan Z, Bai L, Sun J, et al. (2016). Structure of the eukaryoticreplicative CMG helicase suggests a pumpjack motion fortranslocation. Nat Struct Mol Biol 23:217–24.

Zegerman P, Diffley JF. (2007). Phosphorylation of Sld2 andSld3 by cyclin-dependent kinases promotes DNA replica-tion in budding yeast. Nature 445:281–5.

Zellner E, Herrmann T, Schulz C, Grummt F. (2007). Site-spe-cific interaction of the murine pre-replicative complex withorigin DNA: assembly and disassembly during cell cycletransit and differentiation. Nucleic Acids Res 35:6701–13.

Zhang J, Yu L, Wu X, et al. (2010). The interacting domains ofhCdt1 and hMcm6 involved in the chromatin loading ofthe MCM complex in human cells. Cell Cycle 9:4848–57.

Zhang X, Shaw A, Bates PA, et al. (2000). Structure of theAAA ATPase p97. Mol Cell 6:1473–84.

Zhang Z, Hayashi MK, Merkel O, et al. (2002). Structure andfunction of the BAH-containing domain of Orc1p in epi-genetic silencing. EMBO J 21:4600–11.

Zhong W, Feng H, Santiago FE, Kipreos ET. (2003). CUL-4 ubi-quitin ligase maintains genome stability by restrainingDNA-replication licensing. Nature 423:885–9.

Zhou C, Huang SH, Jong AY. (1989). Molecular cloning ofSaccharomyces cerevisiae CDC6 gene. Isolation, identifica-tion, and sequence analysis. J Biol Chem 264:9022–9.

Zou L, Mitchell J, Stillman B. (1997). CDC45, a novel yeastgene that functions with the origin recognition complexand Mcm proteins in initiation of DNA replication. Mol CellBiol 17:553–63.

Zwerschke W, Rottjakob HW, Kuntzel H. (1994). TheSaccharomyces cerevisiae CDC6 gene is transcribed at latemitosis and encodes a ATP/GTPase controlling S phase ini-tiation. J Biol Chem 269:23351–6.

144 M. W. PARKER ET AL.