Annual Reviews - Replication

download Annual Reviews - Replication

of 36

Transcript of Annual Reviews - Replication

  • 8/4/2019 Annual Reviews - Replication

    1/36

    Annu. Rev. Biochem. 2005. 74:283315doi: 10.1146/annurev.biochem.73.011303.073859

    Copyright c 2005 by Annual Reviews. All rights reserved

    CELLULAR DNA REPLICASES: Componentsand Dynamics at the Replication Fork

    Aaron Johnson2 and Mike ODonnell1,21Howard Hughes Medical Institute, 2The Rockefeller University,

    New York City, New York 10021-6399; email: [email protected],

    [email protected]

    Key Words DNA replication, DNA sliding clamps, DNA polymerase, processivityclamp loader, protein-DNA interactions

    Abstract Chromosomal DNA replicases are multicomponent machines that haveevolved clever strategies to perform their function. Although the structure of DNAis elegant in its simplicity, the job of duplicating it is far from simple. At the heartof the replicase machinery is a heteropentameric AAA+ clamp-loading machine thatcouples ATP hydrolysis to load circular clamp proteins onto DNA. The clamps encircleDNA and hold polymerases to the template for processive action. Clamp-loader and

    sliding clamp structures have been solved in both prokaryotic and eukaryotic systems.The heteropentameric clamp loaders are circular oligomers, reflecting the circularshape of their respective clamp substrates. Clamps and clamp loaders also functionin other DNA metabolic processes, including repair, checkpoint mechanisms, and cellcycle progression. Twin polymerases and clamps coordinate their actions with a clamploader and yet other proteins to form a replisome machine that advances the replicationfork.

    CONTENTS

    INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284THE E. coli REPLICASE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

    Pol III Core . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

    The Sliding Clamp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288

    The Complex Clamp Loader . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290

    The Holoenzyme Particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293

    Replisome Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293

    Protein Trafficking on DNA Sliding Clamps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

    THE EUKARYOTIC REPLICASE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296

    Proliferating Cell Nuclear Antigen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298

    The RFC Clamp Loader . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300Eukaryotic DNA Polymerases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

    The Eukaryotic Replisome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304

    Alternate RFC Complexes and Other Roles for Clamp Loaders . . . . . . . . . . . . . . . . 307

    CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307

    0066 4154/05/0707 0283$20 00 283

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    2/36

    284 JOHNSON ODONNELL

    INTRODUCTION

    Chromosomal DNAs are exceedingly large molecules, and the vast repository of

    information they hold must be duplicated with precision [see (1) for overview].The sheer size of these molecules may explain why all cells utilize a clamp loader

    to position a circular sliding clamp on DNA that tethers the DNA polymerases to

    their long substrates for highly processive synthesis (see Figure 1). The eukaryotic

    proliferating cell nuclear antigen (PCNA) and prokaryotic () clamp proteins have

    unrelated sequences, yet they have strikingly similar structures and thus share a

    common ancestor (2, 3). The clamps must be opened and closed around DNA, and

    this job is performed by a multiprotein clamp-loading ATPase [reviewed in (4)].

    Figure 1 Components of cellular replicases. The table lists the three replicase componentsin the well-defined systems of Escherichia coli, eukaryotes, Archaea, and T4 phage. The

    scheme below is a generalized mechanism of replicase action. A multiprotein clamp-loader

    couples ATP binding and hydrolysis to loading of a ring-shaped processivity clamp that is

    then used by the replicative polymerase as a tether to the DNA template.

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    3/36

    CELLULAR REPLICASES 285

    The eukaryotic replication factor C (RFC) and the prokaryotic ( complex) clamp-

    loader subunits are arranged in a circle and are each members of the AAA+ protein

    family [briefly reviewed in (4)].

    The replicases of all cells function within a greater context, involving many otherproteins. For example, two basic activities at a replication fork include a helicase

    to separate the duplex DNA and a primase that produces short RNA primers,

    which are required to initiate DNA synthesis. This larger machinery, frequently

    termed the replisome, is fairly well defined in prokaryotes (1, 58). In E. coli,

    and presumably other bacteria, the leading- and lagging-strand polymerases are

    connected to the clamp loader, which also binds a homohexameric helicase (DnaB

    in E. coli). The helicase acts on the lagging strand and activates the RNA primase

    (DnaG in E. coli). The replicase also makes a specific functional connection to

    single-stranded DNA-binding protein (SSB), which is present at the fork to protectsingle-stranded DNA (ssDNA) and melt hairpins.

    The eukaryotic replisome, in contrast, involves many more proteins and is less

    defined at the current time [reviewed in (911)]. The eukaryotic helicase, primase,

    and SSB are all heterooligomers. The helicase is thought to be the heterohexam-

    eric MCM complex (12), and the primase is the four-subunit DNA polymerase

    /primase that makes a hybrid RNA/DNA primer (13). The heterotrimeric repli-

    cation protein A (RPA) functions as the SSB (14). Several additional eukaryotic

    proteins without prokaryotic counterparts are thought to act during replication ini-

    tiation and fork progression as well (e.g., Cdc45, GINS complex, Dpb11, Sld2,and Sld3) (1517). The functions of most of these components are presently un-

    known. In addition, the leading- and lagging-strand polymerases are thought to be

    different enzymes, Pol (34 subunits) and Pol (4 subunits). It is still unclear

    which strands these polymerases act upon. Finally, many of these proteins are

    regulated by posttranslational modification, including the PCNA clamp which is

    ubiquitinated and sumoylated (1820).

    The clamp and clamp loaders of cellular replicases, prokaryote and eukaryote

    alike, also function in several other processes besides replication. For example,

    PCNA and bind DNA ligase as well as mismatch and excision repair proteins,although the exact role of these interactions is still not entirely clear (2123).

    A weak consensus sequence for proteins that bind PCNA reveals a broad array

    of additional proteins that bind this clamp and are generally involved in repair,

    chromatin structure, or cell cycle control (21). The clamps also function with

    other DNA polymerases (23, 24), probably helping to target them to sites where

    their action is required. Eukaryotes even utilize alternative RFC clamp loaders in

    which one subunit is replaced by a unique protein that presumably specializes the

    complex for the alternative function (2527, 234243). These alternative clamp

    loaders load PCNA clamps at specific target sites or, in one case, load an alternatePCNA-like clamp.

    Because of space limitations, this review focuses on prokaryotic and eukaryotic

    replicases and forgoes a description of archaebacterial, bacteriophage, and viral

    replicases. However, archaebacteria also utilize a similar clamp and clamp-loader

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    4/36

    286 JOHNSON ODONNELL

    strategy, as does the T4 bacteriophage (see the table in Figure 1) [reviewed in

    (7, 28), respectively]. We regret that space limitations prevent discussion of the

    phage T4 replisome in a similar manner; it is an enormously valuable system from

    which many basic early concepts of replisome action emerged, and it continuesto provide fresh insights. A recent review in this series (7) expertly distills the T4

    contributions to date and how they relate to the E. coli system. The replicative

    polymerases of most bacteriophage and viruses do not utilize a clamp and clamp

    loader, but these polymerases generally require one or two accessory factors that

    confer processivity [see T7 phage (7, 29, 30), vaccinia virus (31, 32), herpes virus

    (33, 34), and Pol (35)]. The authors have also refrained from discussing DNA

    replication initiation due to space limitation. Many useful reviews exist on this

    topic [see (1, 28, 36, 37)].

    THE E. coliREPLICASE

    The first cellular replicase to be studied, E. coli DNA polymerase III holoenzyme,

    was isolated as an intact particle from E. coli extracts in the early 1980s, and its

    structure and function serve as a suitable paradigm for its eukaryotic counterpart

    (1). We, therefore, begin this review with the prokaryotic replicase machinery.

    Pol III Core

    DNA polymerase III (Pol III) was originally identified in a mutant strain ofE. coli,

    polA (38). This strain lacked the comparatively strong DNA polymerase I activity,

    unmasking the replicative polymerase. The identity of this activity was likely the

    core polymerase subcomplex. The specific activity of Pol III core is similar to Pol I,

    but as will be described below, Pol III core functions with accessory proteins that

    convert it to an exceedingly efficient enzyme having the highest specific activity

    of any E. coli DNA polymerase (1).

    Pol III core is a 1:1:1 heterotrimer of the polymerase, 3-5 proofreading ex-onuclease, and subunits (39, 40) (see Table 1). The subunit (encoded by dnaE)

    contains the DNA polymerase activity, incorporating 8 nucleotides/second (ntd/s),

    similar to Pol III core (20 ntd/s) (41). The (dnaQ) subunit is the proofreading

    3-5 exonuclease. It is interesting to note that without the processivity of holoen-

    zyme is markedly reduced from >50 kb to about 1.5 kb (42), thereby ensuring that

    the proofreading subunit is present during genome duplication. In contrast, the

    small subunit (holE) has no known function besides a slight stimulation of (43)

    and the holEgene can be deleted with little consequence (44).

    Little structural information is available for Pol III core. The subunit is amember of the C family of DNA polymerases (45). Members of this family are

    present only in bacteria, and it remains the only major family for which no crystal

    structure has been solved. One may presume that it will contain the characteristic

    palm, thumb, and fingers domains and two-metal mechanism of catalysis present

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    5/36

    CELLULAR REPLICASES 287

    TABLE 1 Escherichia coli replisome components and associated functionsa

    in all other polymerases (46). Deletion mutagenesis suggests that the active site islocated in the N-terminal two thirds of, whereas contacts to other holoenzyme

    components are made through the C-terminal region (47, 48).

    Atomic resolution structures of portions of and are available (4951). The

    subunit is composed of two domains (52). The N-terminal domain (186 residues

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    6/36

    288 JOHNSON ODONNELL

    out of 243) contains the exonuclease active site and the binding site. The crystal

    structure of this active fragment has been solved (49), and it is structurally similar

    to other polymerase-associated exonucleases and the exonuclease domain of DNA

    polymerase I. The preferred DNA substrate for is a single-stranded 3

    terminus,and a fully base-paired recessed 3 terminus is a poor substrate for alone (53, 54).

    However, tightly associates with the C terminus of and stimulates activity on

    a recessed 3 terminus, probably by bringing to the primer site. preferentially

    degrades a primer/template with a mismatched recessed 3 terminus. In addition,

    polymerization is inhibited by a mismatch, which provides more time to complete

    the excision.

    The structure of bears a resemblance to a DNA-interacting domain of eukary-

    otic DNA polymerase (50, 51). This structural similarity may underlie the slight

    stimulation of by and increased mutator phenotype of a holE-null mutant in an mutant background (44).

    The Sliding Clamp

    Pol III core by itself is slow, incorporating 20 ntd/s, and weakly processive, ex-

    tending only 110 bases per binding event (41). In fact, no matter how much time

    or enzyme is available, the Pol III core cannot extend a unique primer full circle

    around an M13 ssDNA genome (55, 56).

    To becomean efficient replicase, the core polymerase requires the clamp. Cou-pled to , core becomes exceedingly fast (750 ntd/s) and processive (>50 kb).

    Biochemical studies initially revealed that binds DNA topologically (57), im-

    plying it has a ring shape and encircles the duplex, whereupon it freely slides along

    it. This hypothesis was quickly proven by structure analysis (3) (Figure 2). Core

    polymerase directly associates with (5759), initially occupying about 22 base

    pairs (bp) of primer (60, 61). As core extends DNA, it pulls the clamp along

    behind it.

    The crystal structure of the dimer (3) shows that the two crescent-shaped

    protomers form a ring with a large central channel of35 A in diameter that mayeasily accommodate double-stranded (ds) DNA modeled inside (Figure 2a,b).

    In fact, room exists for one or two layers of water between the DNA and ,

    suggesting may ice skate along the duplex. The center of the ring is lined with

    12 -helices, pairs of which are supported by an outer -sheet. This helix pair and

    sheet motif forms one globular domain, which is repeated six times around the ring,

    creating a sixfold pseudosymmetry. Each monomer consists of three domains

    and dimerizes head-to-tail with another to produce two structurally distinct faces.

    One face has several loops and protruding C termini. This is the face of the ring

    that interacts with other proteins as discussed below (62, 63).The clamp is a tight dimer, and its half-life on DNA is over 1 h at 37C

    (64, 65). The subunit of complex can open by itself, as determined by its

    ability to rapidly remove the clamp from DNA (64, 66). It appears that only one

    binds 2, and it does not dissociate the dimer into monomers, even though it

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    7/36

    CELLULAR REPLICASES 289

    Figure 2 Structure and dynamics of the sliding clamp. (a,b) Crystal structure of the

    dimer with modeled double-stranded DNA through the central channel (courtesy ofD. Jeruzalmi and J. Kuriyan). Panel (b) highlights the two faces of the ring. The C-terminal

    face (right) is implicated in many of the interactions of with other proteins. (c) Superposi-

    tion of one subunit from the dimer structure (purple) and the monomer from the 11crystal structure (yellow), using domain II as a reference. (d) Model of an open clamp made

    by arranging two monomer structures from 11 to create one dimer interface [described in

    (67)]. Panels (c) and (d) adapted from (67), copyright 2001, with permission from Elsevier.

    must destabilize one interface (62, 65, 66). This idea is consistent with the abilityof complex to load that is cross-linked across one dimer interface (66). A

    cocrystal structure of bound to a monomeric mutant of has been solved (67).

    This structure shows two distinct points of contact between and . The contact

    points are located on the opposite ends of the same -helix, which is located in the

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    8/36

    290 JOHNSON ODONNELL

    N-terminal domain of. One end of the helix binds a hydrophobic pocket between

    domains II and III of . The other end appears to push on a loop at the dimer

    interface, leading to a distorted interface that can no longer close. It is proposed

    that getsagripon via the hydrophobic pocket and pushes on the loop to crack theinterface or to hold the interface open. Mutational analysis of these contact points

    has confirmed their roles in clamp-loader interaction and clamp opening (68).

    The - structure explains how the interface is cracked but not how the ring

    opens to accommodate DNA in the central channel. Comparison of the dimer

    and monomer (from the - structure) shows significant rigid body motions

    between the domains, leading to a shallower crescent shape in the monomer (67)

    (seeFigure2c).Thisimpliesthatthe ringisunderspringtensionintheclosedstate

    until disrupts one interface, allowing tension between the domains to relax and

    producing a gap for DNA-strand passage (see Figure 2d). The strong interactionsat the dimer interfaces likely maintain this tension.

    The Complex Clamp Loader

    The E. coli clamp loader is a complex composed of five different subunits in a

    defined stoichiometry: 31

    1 1 1 (69, 70). The complex harnesses the energy

    of ATP binding and hydrolysis to topologically link to a primed DNA, then it

    ejects from DNA, leaving the closed clamp behind (57, 66, 71). It is convenient to

    dissect the clamp loader by the primary function of each subunit (4). The three subunits are the only subunits that bind ATP (72, 73) and have thus been termed

    the motor of the complex. The subunit is called the wrench because it is the

    main clamp-interacting subunit, and it can open the dimer interface by itself.

    The subunit modulates - contact (66). appears to be a rigid protein (74).

    Unlike and ,thedomainsof have more intramolecular interactions and assume

    the same orientation in the structure and within 3. This feature has earned

    the term stator, the stationary part of a machine upon which other parts move.

    In contrast, the three domains of assume different orientations in the trimer.

    The domains of also assume different orientations in - compared to 3

    .The 311 complex is termed the minimal clamp loader, as it is sufficient to

    place on a DNA template (75). The and subunits are not essential for the

    clamp-loading mechanism (76), but links the clamp loader to SSB and primase

    (77, 78), which will be discussed below. serves as a connector to (76) and

    also strengthens the 3 complex (79).

    A broad outline of the clamp-loading mechanism has been determined from

    biochemical studies (4, 66). The nucleotide-free clamp loader has a low affinity

    for the clamp (62). The subunit, which binds tightly to , is likely sequestered

    by the other subunits without ATP present. Upon ATP binding, the complexundergoes a conformational change (62, 71) that allows tight binding to the clamp,

    whereupon opens one dimer interface of the clamp (65). The clamp-clamp loader

    complex has a strong affinity for DNA, particularly a primed template (71, 80).

    The DNA, presumably threaded through the clamp, stimulates ATPase activity in

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    9/36

    CELLULAR REPLICASES 291

    complex, and this allows the ring to close around DNA and the clamp loader to

    eject, thereby recycling it for repeated rounds of clamp loading (66, 81). Hydrolysis

    of ATP leaves the clamp loader in a low-affinity DNA-binding state, presumably

    until ADP dissociates and the complex can be recharged with triphosphatenucleotides (80, 8284).

    The crystal structure of the 3 complex provided deeper insight into clamp-

    loader action (70). This accomplishment was followed two years later by the crystal

    structure of the eukaryotic clamp loader bound to its clamp (85), which has filled

    in many details that will be discussed later. The five subunits ofE. coli 3 are

    arranged in a circular fashion (70) (Figure 3b). The three subunits are adjacent

    to one another, flanked on one side by and the other by , which in turn interact

    with each other to complete the ring. Each subunit consists of three domains, and

    oligomerization is mediated mainly by the C-terminal domain III (Figure 3a,b).The N-terminal domains I and II of all five subunits adopt the chain fold of the

    AAA+ (ATPases associated with a variety of cellular activities) family (Figure 3a)

    [reviewed in (86, 87)]. Although only binds ATP (72), has conserved elements

    of the AAA+ family (88). In contrast, the sequence has diverged, making the

    discovery of its AAA+ fold surprising.

    The ATP sites of complex are located at subunit interfaces and are supple-

    mented with residues from the adjacent subunit (see Figure 3c). The interfacial

    location of ATP sites is typical of AAA+ oligomers and is presumed to promote

    communication among the subunits.

    contributes a key arginine to ATP Site 1of the complex, as seen in the left panel of Figure 3c (red residue) with an

    ATP molecule modeled along the phosphate-binding loop (blue). This arginine is

    located in a conserved serine-arginine-cysteine (SRC) motif that is present in all

    known clamp loaders. Mutation of the Arg, and the homologous residue in ,

    causes a severe defect in clamp-loading and ATPase activity and also disrupts in-

    teraction with DNA ( mutants) or ( mutants) (89, 90). Mutation of the P-loop

    of has more severe consequences that result in complete loss of activity (73).

    These observations highlight the coordination of the ATP cycle with the clamp-

    loading mechanism, suggesting regulation of substrate binding by conformationalchanges in distinct ATP sites during nucleotide binding and hydrolysis.

    The C termini (domains III) of the five subunits of the minimal complex form

    a tight ring, or collar (91), but the N-terminal AAA+ domains (I and II) are more

    loosely associated, with a total lack of contact between and in these domains,

    creating a gap in the N-terminal portion of the ring (70) (Figure 3b). This gap may

    function to allow DNA to pass into positioned under the clamp-loader complex,

    as discussed in the RFC section of this review. Both ATP- and -clamp binding oc-

    cur in these N-terminal domains (67, 92). The loose connections of these domains

    has led to the suggestion that some conformational freedom of the N termini is es-sential for function (93). The structure of3 lacks ATP and thus is in the inactive

    state. Consistent with this, a dimer cannot be docked onto the complex (replac-

    ing with -) without significant clashes, even with the substantial gap between

    and (70). Biophysical experiments in solution indicate that the distance between

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    10/36

    292 JOHNSON ODONNELL

    Figure 3 Crystal structure of the E. coli 311 complex. (a) Isolated subunits from

    the structure of the complex, domains IIII noted [adapted from (87), copyright 2002

    Nature Publishing Group (http://www.nature.com)]. (b) Side view of3 (left). The

    -interacting helix of is marked in yellow. View from the C termini of3 (right)

    [adapted from (70), copyright 2001, with permission from Elsevier]. (c) The ATP sitesof the ring-shaped complex are located at subunit interfaces, as illustrated in the

    cartoon (right) and taken from the structure of ATP Site 1 (left) with ATP (purple)

    modeled against the P-loop (blue). The Ser-Arg-Cys (SRC) motif of , conserved in

    all clamp loaders, points its arginine (red) toward the phosphate of ATP [adapted

    with permission from (89)].

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    11/36

    CELLULAR REPLICASES 293

    and does not change dramatically during clamp association (93). Hence, some

    other conformational change may occur that does not necessitate increasing the

    gap between and . For example, may be pushed downward relative to , thus

    exposing the -binding element of without changing the distance between and. In addition, the subunit also binds (94). Taken together, these observations

    imply that an extensive surface of interaction is formed between the clamp and

    clamp loader. How the clamp loader binds the clamp and DNA is illuminated by

    the RFC-PCNA structure discussed later in this review.

    The Holoenzyme Particle

    On the basis of intracellular Pol III subunit concentrations, a portion of clamp

    loader is associated with the Pol III holoenzyme, but a majority of the clamp

    loader is free in solution (64, 95). The clamp loader associated with the holoenzyme

    contains a different form of the dnaX gene. The subunit is produced through a

    ribosomal frameshift in the dnaXgene, which causes almost immediate termination

    of translation to produce a 47.5-kDa protein (96, 97). The full-length product of

    dnaX is the subunit (71.1 kDa), which contains the sequence plus a unique

    23.6 kDa C-terminal region (c) (Figure 4a). The 23.6 kDa c is comprised of two

    domains, IV and V, that bind DnaB and Pol III core (through ), respectively (98,

    99). The c region is not required for clamp loading but is essential for cell viability

    (100), probably owing to its ability to organize the replisome as discussed below.

    The DnaX protein is present in three copies in the clamp-loader complex, and

    thus various species may assemble in stoichiometries of 3, 21, 12, and 3(69). Each C terminus will recruit one polymerase, and therefore the more

    present in the clamp loader, the more core polymerase molecules it may bind.

    At least two polymerases are required for concurrent synthesis of leading and

    lagging strands (Figure 4b). Because of this requirement, it is thought that the

    E. coli replicase contains two Pol III cores attached to a 21 clamp loader

    and that a specific order of assembly leads to this particle (69, 101, 102), which

    is termed Pol III (or Pol III star). The clamp associates with Pol III in an

    ATP-dependent manner to form the Pol III holoenzyme. The single-copy subunits

    of the clamp loader define an inherent asymmetry, and thus by definition, the

    two cores attached to the two subunits are in somewhat different environments

    (8). The consequence of this asymmetric structure might be minimal because a

    proline-rich segment of separates the clamp-loader and polymerase-interacting

    domains, suggesting a flexible connection. However, DnaB or other holoenzyme

    subunits may hold the cores in defined asymmetric positions (103). It has been

    proposed that the asymmetric structure imposes distinct properties onto the two

    polymerases, modeling their behavior to fit the different needs of replicating the

    leading and lagging strands (8, 102, 104109).

    Replisome Dynamics

    The subunits of Pol III holoenzyme not only connect two core polymerases

    to the central clamp loader, but also connect the replicase to the DnaB helicase

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    12/36

    294 JOHNSON ODONNELL

    Figure 4 Organization and dynamics of the E. coli replisome. (a) The subunit of DNA

    polymerase III holoenzyme is comprised of the clamp-loader domains IIII ( sequence)

    and the replisome organization domains IV and V that bind the DnaB helicase and Pol III

    core, respectively. (b) Polymerase cycling at the replication fork. As the replisome advances,the clamp loader loads a clamp on an RNA primer (pink) synthesized by DnaG (upper

    right). When the lagging-strand polymerase replicates to a nick, it dissociates from DNA and

    (lower right) and cycles to the newly loaded clamp (lower left).

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    13/36

    CELLULAR REPLICASES 295

    (see Figure 4b). The homohexameric DnaB encircles the lagging strand, and when

    coupled to DNA synthesis through , its unwinding rate increases from 35 bp/s

    to near holoenzyme speed (98, 99, 103, 110112).

    As the replisome advances, the polymerase on the leading strand simply extendsDNA in a continuous fashion. Presumed impediments to leading strand extension

    include sites of DNA damage and the consequent collapse of the fork. Repair and

    restart of synthesis is an elaborate process detailed in recent reviews (113, 114). Pol

    III holoenzyme has been implicated in mediating the restart of DNA replication

    after fork stalling (115).

    Lagging-strand replication is a discontinuous process of fits and starts that re-

    peats in a cycle time of 13 s. An overview of the lagging-strand cycle is illustrated

    in Figure 4b. Each Okazaki fragment is initiated by primase, which synthesizes

    an RNA primer of about 1012 nucleotides (116, 117). Primase action requiresinteraction with DnaB, which involves a C-terminal region of primase (118, 119).

    Primase extends the RNA in the opposite direction of helicase unwinding and is

    presumed to separate from DnaB, which may account for its observed distributive

    action (120). Primase remains attached to the RNA primed site through its inter-

    action with SSB (121123). Although primase eventually dissociates, release of

    primase is accelerated by the subunit of the clamp loader, which binds SSB in

    a competitive fashion, recruiting the clamp loader to the DNA template to com-

    pete with primase (124). The clamp loader then places onto the primer for the

    lagging-strand polymerase.As the lagging polymerase extends a fragment, a loop is generated because it

    is connected to the leading polymerase (via the clamp loader), yet extends DNA

    in the opposite direction (see top left diagram), as originally proposed (125) and

    recently confirmed (126) in the T4 system. The 13 kb Okazaki fragment will be

    completed within a few seconds (Figure 4b, top right diagram), and at this point,

    the core must rapidly release from DNA to start the next fragment (bottom right

    diagram). The highly processive Pol III requires a specific mechanism for this

    release step, which disengages core from , leaving the clamp behind on the

    finished fragment. The release step occurs only at a nick, thus ensuring completionof the fragment, and requires the subunit (127130). The lagging-strand core

    is now free to bind a new clamp placed on the next RNA primer by the clamp

    loader (bottom left diagram).

    Replication fork progression has been studied in E. coli, using rolling-circle

    DNA templates (103, 107, 116, 120, 131133). That clamps accumulate on

    the lagging strand and that the single clamp loader can rapidly load clamps on

    DNA repeatedly during fork progression have been demonstrated using this system

    (103). The accumulated clamps can be recycled by the unloading action of

    complex and also by the fivefold excess of free subunit in the E. coli cell (64).

    Protein Trafficking on DNA Sliding Clamps

    Many different proteins function with sliding clamps. An understanding of the

    way proteins coordinate their traffic flow on clamps is now emerging. To illustrate

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    14/36

    296 JOHNSON ODONNELL

    how proteins may switch positions to utilize the clamp, we briefly describe three

    important switches on that occur during the progression of theE. coli replication

    fork.

    The / clamp loader and Pol III core are known to bind the same face of inessentially the same spot, and therefore these two factors compete for the clamp

    (63, 134). Yet both the core and / complex must function with at the start of

    each DNA chain. This protein trafficking event is facilitated by the fact that ATP

    hydrolysis ejects the clamp loader from (Figure 5a) (71). In addition, the clamp

    loader remains in a reduced activity state for a short interval, presumably owing to

    slow ADP release (80, 83). The core polymerase is then free to bind the abandoned

    clamp and, in fact, binds tighter on DNA than off (63).

    Switching on is also thought to occur when the leading polymerase stalls.

    In this case it is thought that a bypass polymerase, either Pol IV or Pol V, takespossession of the clamp to extend DNA through the stall site. Recent structural

    analysis suggests that Pol IV can associate in two ways with (135, 136). Pol IV

    binds the edge of the ring and one hydrophobic pocket, but Pol IV is angled

    off the DNA (135). It is possible that in this binding mode Pol IV may bind to

    simultaneously with Pol III (see Figure 5b). Upon stalling of Pol III, Pol IV

    must break its interaction with the side of the ring and swing down to the DNA,

    presumably maintaining its hold on the hydrophobic pocket of . This action

    would displace Pol III from DNA and perhaps disrupts the Pol III- contact as

    well. Pol IV is distributive, even with , allowing Pol III to regain the clampafter the lesion is bypassed. Recent evidence with Pol III, Pol V, and suggests

    that a similar trade-off may occur (137). At the end of an Okazaki fragment, the

    normally highly processive Pol III rapidly dissociates from (127130), freeing

    the polymerase to extend the next Okazaki fragment. This trafficking event is

    mediated by the c portion of.The subunit binds via the extreme C-terminal

    residues (134). c also binds the C terminus and disrupts core- interaction.

    However, c binds ssDNA, andthis prevents from binding the C-terminal residues

    of . Hence, so long as there is ssDNA template, c is turned off, and core

    functions with . But when all available ssDNA is converted to duplex, c turnson and separates core from (see Figure 5c).These switch processes explain how

    interactions with the clamp are modulated by ATP or DNA structure to promote

    the trafficking of different proteins on sliding clamps to ensure progression of

    replication.

    THE EUKARYOTIC REPLICASE

    Although the complexity of architecture, interaction, and regulation of DNA repli-cation is far greater in eukaryotes than in bacteria, the core replicase components

    are structurally and functionally more similar than different. However, beyond this

    basic machinery lies a much larger network of proteins required for propagation

    of the replication fork and regulation of its advance as well as coordinating its

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    15/36

    CELLULAR REPLICASES 297

    Figure 5 Three examples of protein trafficking on sliding clamps. Interactions with

    the clamp fromE. coli. (a)The complex clamploader associates tightly with when

    bound to ATP. DNA triggers ATP hydrolysis, resulting in low affinity for and DNA.

    (b) When Pol III, the replicative polymerase, encounters a lesion in the DNA template,

    it stalls, unable to overcome its inherent fidelity to incorporate opposite a damaged base.

    Stalling allows an error-prone polymerase, such as Pol IV (red) passively traveling on

    , an opportunity to trade places with Pol III on to replicate past the lesion. [Adapted

    with permission from (135).] (c) Pol III maintains a tight grip on via the polymerase

    C terminus. However, when it completely replicates its substrate DNA, the polymerase

    must release from to recycle to the next primed site. The subunit modulates this

    interaction, binding the polymerase C tail only when no more single-stranded template

    is present. This severs the connection between the polymerase and the clamp [adapted

    with permission from (134), copyright 2003, National Academy of Sciences, U.S.A.].

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    16/36

    298 JOHNSON ODONNELL

    activity with other DNA metabolic machineries. Many of these details are only

    now coming into focus.

    Proliferating Cell Nuclear AntigenPCNA derives its name from the early finding that the protein is abundant in pro-

    liferating cells (138). PCNA was shown to be directly involved in DNA replication

    through its ability to stimulate DNA polymerase in replicating long stretches of

    primed DNA (139141). This characteristic suggested it may be analogous to the

    E. coli subunit, although neither were known to be sliding clamps at the time.

    Subsequent work showed the importance of PCNA for simian virus 40 (SV40)

    DNA replication and its enhanced activity in the presence of RFC and ATP (142).

    As a monomeric unit, PCNA is about two-thirds the size of, and accordingly,

    each protomer contains two structurally similar domains instead of the three do-

    mains found in (2). The chain folds of the two PCNA domains are the same

    as those found in the domains of . PCNA and form very similar ring-shaped

    structures, except PCNA must trimerize to form a six-domain ring (Figure 6a).

    Like , the PCNA ring is quite stably attached to DNA (t1/2 = 24 min) (143).

    The PCNA protomers are also arranged head-to-tail to create two distinct faces

    of the ring, mirroring . As in the E. coli system, the eukaryotic clamp loader and

    polymerase compete for binding the same face of the PCNA ring, the face from

    which the C termini project (144, 145).

    A variety of proteins involved in DNA repair and cell cycle control interact

    with PCNA, but this topic is covered elsewhere [reviewed in (21, 23)]. Relevant

    to the current review, a weak PCNA-binding consensus sequence has emerged:

    Q-x-x-h-X-X-a-a, where x = any residue; h = L,I,M; and a = F,Y (21, 23). The

    crystal structure of human PCNA in complex with a peptide derived from the cell

    cycle regulator p21WAF1/Cip1 was the first to demonstrate that these clamp-binding

    Figure 6 Structures of the eukaryotic clamp and clamp loader from Saccharomycescerevisiae. (a) View of the C-terminal face of PCNA. The ring-shaped PCNA is a head-

    to-tail trimer of a two-domain monomer. The sixfold pseudosymmetry of the clamp

    is evident in PCNA as well. (b) The structure of replication factor C (RFC) bound to

    PCNA reveals the structural similarity between RFC and complex. RFC binds to the

    C-terminal face of PCNA. (ce) The RFC subunits are arranged in a helix that tracks

    the minor groove of B-form DNA modeled through the PCNA ring. (d) In this cartoon,

    the 5 terminus of a recessed primer template is positioned to exit the central channel of

    the clamp and clamp loader through the gap between RFC1 and RFC5. ( e) N-terminal

    regions of the five RFC subunits and the PCNA ring from the RFC-PCNA structure.Two conserved helices in each RFC subunit (yellow) are in position to interact with

    DNA (orange/green) that passes through the central channel of PCNA (gray)withthe5

    terminus (green spheres) exiting between RFC1 and RFC5. [Adapted with permission

    from (85), copyright 2004 Nature Publishing Group (http://www.nature.com).]

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    17/36

    CELLULAR REPLICASES 299

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    18/36

    300 JOHNSON ODONNELL

    proteins interact with PCNA at a hydrophobic pocket located between the two

    domains of a protomer (146). The finding generalizes to the position at which

    DNA polymerase binds the T4 gp45 clamp (147149). Also, the subunit of Pol

    III core and subunit of complex bind E. coli at a spot between domains IIand III (24, 67, 150).

    The RFC Clamp Loader

    The eukaryotic clamp loader was first isolated as a required component of the

    in vitro SV40 genome replication system (151) [reviewed in (152)]. The activity

    was originally named replication factor C (RFC) (51) or activator-1 (153). RFC

    is a DNA-dependent ATPase that functions with PCNA to confer processivity on

    DNA polymerase (142) [reviewed in (154, 155)]. The similarity between RFCandE. coli complex was apparent early on. RFC consists of five different proteins

    that are homologous to each other and to and ofE. coli complex (156, 157).

    Common sequence motifs among these clamp-loader AAA+ proteins have been

    termed RFC boxes.

    In overview, RFC acts similarly to complex. In an ATP-dependent reaction,

    RFC loads PCNA onto a recessed 3 primer/template junction and then dissociates,

    allowing PCNA to function with Pol (158, 159). The RFC subunits in S. cere-

    visiae are referred to as RFC15 (157) (see Table 2 for human RFC nomenclature).

    RFC25 share a similar molecular weight and three-domain architecture charac-teristic of the E. coli subunit (85, 160). RFC1 also contains these three domains,

    along with sizable N- and C-terminal extensions (161). The N-terminal region

    (residues 1275 in S. cerevisiae) has clear homology to DNA ligases, although

    there is no evidence of ligase activity. The removal of the RFC1 N terminus re-

    sults in sensitivity to DNA-damaging agents (162), but this region is not necessary

    for in vitro clamp loading (163) or cell viability (162). The C-terminal domain

    (residues 660861 in S. cerevisiae) has not yet been characterized genetically or

    biochemically.

    The five-subunit composition of RFC bears a close similarity to the minimal E. coli 3

    complex (85). On the basis of subunit interactions and sequence

    similarity to complex subunits, the subunit arrangement of the RFC pentamer

    was proposed (4, 164) and has since been proven by structure analysis (85). RFC1

    shares characteristics of the subunit in its conserved clamp-interacting residues

    and position in the pentamer (163, 165). RFC24 are considered -like in their

    ability to form a trimeric ATPase subassembly (166). RFC5 is in the position of,

    and like it contains an SRC motif but lacks a consensus phosphate-binding loop

    (P-loop). Like the 3 complex, the ATP sites of RFC are located at subunit

    interfaces. However, unlike 3 , RFC contains four competent ATPase sitesbecause RFC1 also binds ATP where does not. In fact, RFC5 also binds a

    nucleotide in the crystal structure.

    During the RFC ATPase cycle in S. cerevisiae, the complex initially binds

    two ATP, then a third upon PCNA binding, and a fourth when it locates the

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    19/36

    CELLULAR REPLICASES 301

    TABLE 2 Eukaryotic replisome components

    Replisome

    component

    Saccharomycescerevisiae

    (kDa) Human (kDa)

    Function and remarks[Schizosaccharomyces pombe

    name (S.p.)]

    RFCa RFC (277.7)a RFC (314.9)a Pentameric clamp loadera

    RFC1 (94.9) p140 (128.2) Binds ATP; phosphorylated

    RFC2 (39.7) p37 (39.2) Binds ATP

    RFC3 (38.2) p36 (40.6) Binds ATP

    RFC4 (36.1) p40 (39.7) Binds ATP

    RFC5 (39.9) p38 (38.5) Binds ATP or ADP

    PCNA PCNA (28.9) PCNA (28.7) 87 kDa homotrimeric processivity

    sliding clampa

    Pol a Pol (220.2)a Pol (238.7)a Replicative DNA polymerasea

    Pol3 (124.6) p125 (123.6) DNA polymerase, 3-5 exonuclease,

    binds PCNA; subunit A (S.p. Pol3)

    Pol31 (55.3) p50 (51.3) Structural subunit; subunit B (S.p. Cdc1)

    Pol32 (40.3) p66 (51.4) Binds PCNA; subunit C (S.p. Cdc27);

    binds Pol large subunit

    p12 (12.4) Structural, stimulates processivity; subunit

    D (S.p. Cdm1)

    Pol a Pol (378.7)a Pol (350.3)a Replicative DNA polymerasea

    Pol2 (255.7) p261 (261.5) DNA polymerase, 3

    -5

    exonuclease(S.p. Pol2/cdc20)

    Dpb2 (78.3) p59 (59.5) Binds polymerase subunit (S.p. Dpb2)

    Dpb3 (22.7) p17 (17.0) Binds Dpb4

    Dpb4 (22.0) p12 (12.3) Present in ISW2/yCHRAC chromatin-

    remodeling complex (S.p. Dpb4)

    Pol a Pol (355.6)a Pol (340.6)a DNA polymerase/primasea

    Pol1 (166.8) p180 (165.9) DNA polymerase

    Pol12 (78.8) p68 (66.0) Structural subunit

    Pri2 (62.3) p55 (58.8) Interacts tightly with p48

    Pri1 (47.7) p48 (49.9) RNA primase catalytic subunit

    MCMa MCM (605.6)a MCM (535)a Putative 3-5 replicative helicasea

    Mcm2 (98.8) Mcm2 (91.5) Phosphorylated by Dbf4-dependent kinase

    Mcm3 (107.5) Mcm3 (91.0) Ubiquitinated, acetylated

    Mcm4 (105.0) Mcm4 (96.6) Helicase with MCM6,7; phosphorylated

    by CDK; aka Cdc54

    Mcm5 (86.4) Mcm5 (82.3) Aka Cdc46; Bob1 is a mutant form

    Mcm6 (113.0) Mcm6 (92.3) Helicase with MCM4,7

    Mcm7 (94.9) Mcm7 (81.3) Helicase with MCM4,6; ubiquitinated

    RPAa

    RPA (114)a

    RPA (100.5)a

    Single-stranded DNA-binding proteina

    RPA70 (70.3) RPA70 (70.3) Binds DNA, stimulates Pol

    RPA30 (29.9) RPA30 (29) Binds RPA70 and 14, phosphorylated

    RPA14 (13.8) RPA14 (13.5) Binds RPA30

    aInformation concerns a protein complex.

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    20/36

    302 JOHNSON ODONNELL

    primer/template DNA, which triggers ATP hydrolysis, causing RFC to eject and

    leave the closed PCNA ring on DNA (167). This mechanism excludes the route

    of RFC first encountering DNA and subsequently recruiting PCNA. ATP site mu-

    tational studies have demonstrated that, in general, disrupting any one of the fourconsensus ATP sites has a significant effect on the activity of RFC (168170),

    although ATP binding may suffice to rescue this defect in at least one ATP site (in

    RFC4 for S. cerevisiae). These single P-loop mutants still bind PCNA readily but

    are deficient in DNA binding (170).

    A crystal structure of an RFC-PCNA complex has been solved (85). To promote

    stability of this nucleotide-dependent interaction by preventing hydrolysis, ATPS

    was used, and RFC was produced with R to Q mutations in the SRC motifs. The

    RFC-ATPS-PCNA crystal structure has provided a detailed view of how a clamp

    loader interacts with its cognate clamp (see Figure 6b). In addition, the structure hasrevealed how DNA binds to a clamp loader. The main PCNA contacts occur through

    RFC1 and RFC3, which bind to the hydrophobic pocket between the domains of

    two different PCNA protomers. RFC1 interacts extensively with PCNA, whereas

    RFC3 seems to be only partially engaged with PCNA and the ring is closed. In

    the RFC-ATPS-PCNA structure, RFC2 and RFC5 do not bind PCNA at all. This

    suboptimal interaction of RFC with PCNA may explain why the clamp remains

    closed and only slightly perturbed from its unbound structure. Alternatively, the

    closed PCNA may be due to crystal packing forces, instability of an open-ring

    complex, or the arginine to glutamine mutation in the four SRC motifs. Althoughthis mutation should ensure that no ATPS becomes hydrolyzed, it may have

    prevented or perturbed some necessary RFC subunit-PCNA interaction needed for

    clamp opening.

    The way RFC binds DNA is suggested by the helical arrangement of RFC sub-

    units in the structure, with the helical axis passing through the central channel of

    the closed PCNA (see Figure 6c,e). The helix begins at RFC1, the subunit that is

    fully engaged with the hydrophobic pocket on one PCNA protomer. Adjacent to

    RFC1 is RFC4, displaced by the helical operator of 61 rotation and 5.5 A trans-

    lation. Overall, the helical operations that relate all five subunits have a pitch of5.6 A per 60 rotation, ending with RFC5, which lies 25 A above PCNA and is

    significantly separated from the ATPase domain of RFC1. This right-handed helix

    and pitch mimics that of duplex B-form DNA. Furthermore, there is sufficient

    space in the center of RFC to model a DNA duplex, which passes right through

    the center of PCNA. Each RFC subunit has two -helices that are oriented so their

    positive dipole tracks the minor groove phosphate backbone of DNA modeled into

    the structure (Figure 6e). Several basic residues on these helices are conserved in

    E. coli and eukaryotic clamp loaders, consistent with the idea that DNA binds

    within this central area of RFC. Modeling a 3

    -recessed, primed template from an-other crystal structure into the center of the ring shows that the specific recognition

    of a primer/template by RFC might occur by a simple clash of the primed-template

    junction with the domain III cap of the RFC subunits. A stiff duplex DNA could

    not proceed through the RFC structure because it would hit the cap and could

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    21/36

    CELLULAR REPLICASES 303

    not bend to exit out the side of RFC. The flexible ssDNA of a primed junction

    may bend to exit through the gap between RFC1 and 5, perhaps assisted by a

    positive patch on the RFC1 surface (see Figure 6e). This hypothesis of specific

    primer/template recognition is in agreement with recent biophysical evidence ofRFC/DNA association (171, 172).

    Eukaryotic DNA Polymerases

    Many DNA polymerases are now known to function in the eukaryotic cell. These

    polymerases all share a common catalytic mechanism [reviewed in (11)], but most

    serve a specific function outside of basic genome duplication. DNA polymerases

    , , and are the established replicative polymerases and are thought to function

    together at the replication fork to copy genomic DNA in a semi-discontinuousmanner. Pol , , and are all members of the B-family of DNA polymerases (45).

    DNA POLYMERASE DNA polymerase (Pol ) was initially thought to be the

    main replicase before Pol was discovered. Pol is unique in its ability to initiate

    DNA synthesis by first synthesizing its own12-nucleotide RNA primer and then

    extending it with about 20 bases of DNA (173, 174) [reviewed in (13)]. Pol is

    now thought to be the eukaryotic primase, making a hybrid RNA/DNA primer,

    followed by a polymerase switch allowing the replicase to take over elongation

    (175). The switch is mediated by RFC, which displaces Pol from the primer viacompetition for RPA (176, 177).

    Pol consists of four subunits (13). The DNA polymerase activity is found in

    the largest subunit (p180), and primase activity is located in the smallest subunit

    (p48). The exact functions of the middle two subunits are not clear, but all four

    subunits are present in Pol isolated from yeast, human,Xenopus,andDrosophila.

    DNA POLYMERASE DNA polymerase (Pol ) in fission yeast, human, and other

    eukaryotic organisms is composed of four essential subunits (178, 179). Interest-

    ingly, in S. cerevisiae, Pol has only three subunits, and no apparent homologueexists for the fourth subunit (180). Furthermore, the third subunit can be deleted

    from budding yeast, although cell growth is compromised (181).

    A unified subunit nomenclature for Pol has recently been proposed (182). On

    the basis ofSchizosaccharomyces pombe Pol subunits and their homologues in

    other eukaryotes, the subunits have been renamed AD for the Pol3(A), Cdc1(B),

    Cdc27(C), andCdm1(D) polypeptides. Much debate has centered on the possibility

    of the Pol complex self-associating into a dimer. The most recent evidence in

    multiple systems indicates that Pol complex contains only one copy of each

    subunit across a wide range of concentrations and has an elongated shape thatresulted in the earlier confusion over whether it was a dimeric polymerase particle

    (183, 184). Pol subcomplexes can also be isolated as a core A/B dimer (180,

    185). A zinc-finger module in subunit A interacts with subunit B, which acts as a

    bridge to subunits C and D (when present) (183, 186, 187). The polymerase and

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    22/36

    304 JOHNSON ODONNELL

    3-5 exonuclease activities of Pol are present in the large A subunit. The Pol

    polymerase extends DNA with high fidelity, but the exonuclease seems surprisingly

    ineffective in vitro in S. pombe (188) and may be most important when Pol is

    clamped onto DNA by PCNA (189, 190).Pol associates with PCNA via interactions with at least two of its subunits.

    The clamp is positioned behind the polymerase (191), preventing polymerase

    dissociation as it extends a primer (192). The strongest interaction is between

    PCNA and subunit C (193). Subunit C contains the consensus PCNA interaction

    motif, and studies in yeast show that this PCNA-interacting subunit stimulates Pol

    (183, 194). Surprisingly, stimulation is not dependent upon the PCNA-interacting

    motif but mainly on the domain involved in connection to the A/B complex. The

    A/B heterodimer alone can also associate with PCNA, presumably through subunit

    A (180). The stimulation of subunit C on Pol activity might be due to proteincomplex stabilization rather than catalytic enhancement. Even the small D subunit

    from the human four-subunit Pol complex significantly stimulated the A/B/C

    subcomplex in assays with RFC and PCNA (179).

    DNA POLYMERASE (POL ) Studies showing Pol is essential in yeast placed it at

    the replication fork (195). Indeed, chromatin immunoprecipitation (ChIP) assays

    in yeast indicate that Pol is located at origins prior to S phase (196) and moves

    away from origins upon releasing an S phase block (197), consistent with a role

    in chromosome replication. However, there are some conflicting genetic studieson whether the intrinsic DNA polymerase activity is required for replication or

    whether the protein may instead serve another role, either as a DNA sensor or

    checkpoint protein (198200), or perhaps it holds together other proteins that are

    essential in the replisomal particle (201). Although further work will be required to

    fully understand the exact role of Pol , it is widely believed to be directly involved

    in DNA synthesis at the replication fork.

    A recent report on Pol concludes that it is a heterotetramer with a stoichiometry

    of 1:1:1:1, presumably in all eukaryotes (202). The DNA polymerase and 3 -5

    exonuclease of Pol reside in the largest subunit and appear to have a higherfidelity than Pol /PCNA (203). The third-largest subunit, Dpb4, is also a member

    of a complex that appears to be involved in chromatin remodeling (204) (A. Tackett

    and B. Chait, personal communication). Pol activity does not absolutely require

    PCNA, but PCNA stimulation increases as the ionic strength is raised (205207).

    The PCNA interaction motif on the large subunit of Pol is not essential for cell

    viability, but mutational analysis of this sequence suggests a role in DNA repair

    (208).

    The Eukaryotic Replisome

    The essential nature of both Pol and Pol and the fact that they both function

    with PCNA as monomeric polymerase particles have led to renewed proposals

    that they function together to replicate the leading and lagging strands. Xenopus

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    23/36

    CELLULAR REPLICASES 305

    extracts have shown that depletion of either polymerase results in a dramatic drop

    in replication (209, 210). Depletion of Pol gave rise to many short nascent

    strands that may have been improperly elongated Okazaki fragments, whereas

    Pol depletion merely slowed down the replication machinery. Viability of Pol-deleted strains of yeast suggests that Pol is not essential for cell viability (198,

    201, 211). However, an inactive point mutant of Pol resulted in cell death (212).

    Onepossible explanation for this apparent contradiction is that another polymerase,

    presumably Pol , can carry out DNA replication in the complete absence of Pol

    but that the Pol point mutant acted as a dominant negative to stop the fork.

    Intensive studies of in vitro replication from the SV40 origin have provided

    great insight into eukaryotic fork function, but the SV40 T antigen fills many roles

    ordinarily performed by host proteins, including helicase function (213). Study

    of polymerases in this system show that Pol and Pol , and even Pol aloneunder some conditions, are sufficient for replication. However, the small size of

    the genome and use of T antigen may minimize the requirements for replication

    [reviewed in (152)]. Reconstitution of a eukaryotic replisome on a rolling-circle

    template may greatly facilitate our understanding of the roles of Pol and Pol

    and which strand(s) they operate on.

    The 10-fold-smaller size of eukaryotic lagging-strand fragments (200 bp)

    compared to E. coli is counterbalanced by the 10-fold-slower rate of fork move-

    ment. Stoichiometric use of PCNA clamps during lagging-strand replication has

    not been demonstrated but is presumed to occur as it does with the E. coli clamp. That PCNA clamps are left behind on lagging-strand fragments is implied

    by the fact that PCNA interacts with and, in some cases, stimulates the factors

    necessary for Okazaki fragment maturation (214216). There is also an interest-

    ing observation that PCNA-DNA complexes may persist through mitosis, marking

    chromosomes for epigenetic inheritance (217, 218).

    The eukaryotic replisome factors that contain helicase, primase, and SSB are

    each composed of multi-protein assemblies in eukaryotes [reviewed in (9)]. This

    complexity is in contrast to the single-subunit factors in E. coli. For example, even

    the SSB in eukaryotes, termed RPA, is composed of three subunits in a 1:1:1 het-erotrimer [reviewed in (14)]. It is widely believed that the eukaryotic helicase is the

    heterohexameric MCM2-7 complex [reviewed in (12)], although this conclusion

    is not yet firm. Like E. coli DnaB, the hexameric MCM complex is ring shaped,

    but each MCM subunit is a different polypeptide. Subunit arrangements for the

    MCM2-7 complex have been proposed (219, 220). Helicase activity has been ob-

    served only for the MCM4/6/7 subcomplex (221223), yet all six MCM genes are

    essential in a variety of systems (224) [also, see references within (12)]. These

    findings have led to the suggestion that one or more of the MCM2,3,5 subunits

    act as regulators (219, 222, 225). Consistent with this view, MCM2 inhibits theMCM4/6/7 helicase. The MCM complex is also a target of phosphorylation and

    ubiquitination and is thought to require activation for helicase action after assembly

    on DNA. MCM subunits are AAA+ proteins and thus are thought to have an evo-

    lutionary origin distinct from DnaB, which is constructed from the RecA module.

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    24/36

    306 JOHNSON ODONNELL

    Figure 7 Hypothetical arrangement of proteins at the eukaryotic replication fork. The

    hexameric MCM complex encircles the leading strand. In this cartoon, Pol is placed

    on the leading strand and Pol on the lagging, with RFC bridging the two polymerases

    and helicase. Pol /primase action places it on the lagging strand along with RPA bound

    to the looping single-stranded DNA. Other factors involved in replication and known

    to bind certain proteins at the replication fork include Cdc45, Sld2, Sld3, Dpb11, and

    the heterotetrameric GINS complex.

    Furthermore, these helicases translocate on DNA with opposite polarities (221

    223), thereby placing the MCMs on the leading strand (see Figure 7). However,

    like DnaB, the MCMs have been shown to be capable of encircling two DNA

    strands (223, 226), and evidence that they may form a double hexamer exists in

    both Archaea (227, 228) and S. pombe (229) systems. One line of evidence for

    this comes from an archaeal single-gene MCM that produces a circular double

    hexamer with helicase activity (227).MCM helicase activity is rather weak, somewhat reminiscent of the relatively

    weak helicase activity ofE. coli DnaB. As described earlier, DnaB becomes highly

    active when coupled to Pol III holoenzyme, and this coupling occurs through the

    subunit of the clamp loader (111). It seems likely that a similar arrangement may

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    25/36

    CELLULAR REPLICASES 307

    exist in eukaryotes. In the scheme of Figure 7, RFC is proposed to act as a scaffold

    similar to the / complex in E. coli, bridging the two polymerases and coupling

    them to the helicase. The proposed connections are simply inferred because only

    scant evidence exists for an RFC-Pol interaction (176, 177), and none as yetexists for RFC binding either Pol or the MCM complex. However, evidence

    exists for the presence of many other protein actors required for DNA synthesis

    and presumed to act at the eukaryotic replication fork. Although the individual

    functions of these factors are largely unknown, protein interaction studies indicate

    a network as illustrated schematically in Figure 7. Cdc45 has been known for some

    time to be required for replication, and interaction between Cdc45 and MCMs has

    been documented (230, 231). Newer actors include Sld3, which binds Cdc45 (17),

    and the heterotetramer GINS complex, which appears to have a ring shape (15,

    232). The GINS complex also appears to bind Pol , and similar to Sld3, assemblesat origins just prior to DNA synthesis (15). Dpb11 is thought to bind both Pol

    and Pol (233), and it also forms a complex with Sld2 (16). Biochemical study

    of these various factors, alone and in combinations, will be required to understand

    their individual roles in chromosome replication and to determine whether they all

    function together at each replication fork.

    Alternate RFC Complexes and Other Roles for Clamp Loaders

    Sliding clamps are used in a variety of DNA metabolic processes (21, 23), and one

    may presume that their respective clamp loader is also involved in most of these

    processes. In addition, the subunit composition of the eukaryotic clamp loader is

    altered to perform novel functions in DNA metabolism. This alteration involves

    the use of an alternate clamp-loader subunit, in place of RFC1. In fission yeast

    and human, the Rad17 subunit (Rad24 in S. cerevisiae) replaces p140 (RFC1) in

    complex with RFC25 (25, 234). The Rad17-RFC complex is involved in the DNA

    damage checkpoint response, along with a novel PCNA-like sliding clamp formed

    from the trimer Rad9/Rad1/Hus1 (Ddc1/Rad17/Mec3 in S. cerevisiae) [reviewed

    in (235)]. This clamp loader loads the 911 clamp onto primed DNA in an RPA-

    stimulated reaction (236239) and does so with opposite polarity to RFC (238).

    The function of the 911 clamp is not clear, but it presumably recruits other factors

    when loaded on DNA. There may be a 3-5 exonuclease activity in Rad9 or Hus1,

    thus providing a biochemical activity for the ring itself (240). RFC1 can also be

    replaced by Ctf18/Chl12 (26, 241) or Elg1 (27, 242, 243), which are involved in

    cohesion and genome stability, respectively, although the specific roles of these

    complexes are not understood.

    CONCLUDING REMARKS

    Each passing year brings significant advances in our understanding of replica-

    tion fork mechanisms. However, for each question answered, 10 more crop up.

    Even in the relatively well-defined and intensively studied prokaryotic system

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    26/36

    308 JOHNSON ODONNELL

    there remain numerous questions. We still do not know how the proteins are truly

    arranged at the replication fork. How are the two polymerases oriented, do they

    face the same direction? Does the helicase encircle one or more strands? Does

    it act as a double hexamer? How is the helicase positioned on the replicase?How is the length of Okazaki fragments measured and regulated? What hap-

    pens when the replisome encounters a lesion on the leading or lagging strand?

    Does the helicase continue to unwind DNA, and if so, how far does it go be-

    fore it stops? To advance past a lesion, E. coli mounts the SOS response and

    gets past the lesion by recombinative repair processes, collectively termed repli-

    cation restart [reviewed in (113, 114)]. How does the replication machinery

    coordinate its action with those of repair and recombination? How does the repli-

    some deal with transcribing RNA polymerase and other proteins tightly bound to

    DNA?The sliding clamps in both prokaryotes and eukaryotes interact with many dif-

    ferent DNA polymerases and repair proteins. For example, and PCNA both

    interact with mismatch repair and excision repair proteins. What is the role of the

    clamps in these processes? The clamps also bind a variety of specialized DNA

    polymerases such as those of the Y-family, which have low fidelity but can by-

    pass certain lesions. How do these polymerases trade places with the replicase at

    the right time and place to bypass lesions? How is the use of the clamp by these

    low-fidelity enzymes restricted and handed back to the high-fidelity replicase af-

    ter a lesion is bypassed? PCNA appears to be modified by ubiquitination and/orsumoylation to assist this process. How do these modifications control the traf-

    ficking of different polymerases on PCNA? Eukaryotes use a variety of alternate

    clamp loaders in which RFC1 is replaced by another protein. These RFC1 sub-

    stitutes appear to be involved in the DNA damage response, chromatin cohesion,

    and genome stability. How do these alternate clamp loaders function and do they

    interface with the normal replication machinery? Clearly much work remains to

    be done to address these important issues.

    Knowledge of the eukaryotic replisome composition is rapidly becoming a

    complex challenge. How many more factors are there? How do they connect andwhat are their individual functions? Do all these proteins function within the same

    replisome or are there various types of eukaryotic replisomes, specialized for

    different regions of the chromosome or for different times in S phase? How is

    replisome assembly and function regulated at the start of S phase? How does this

    tie in with growth factors that work at the cell surface, and the complex kinase

    and cell cycle control networks that reach to the nucleus? How do DNA damage

    checkpoint and other checkpoint mechanisms exert their influence during ongoing

    S-phase events?

    The authors hope this review highlighted several important advances in ourknowledge of replicase structure and function. However, these concluding remarks

    underscore how much is yet to be learned and perhaps provide some sense of how

    far we have yet to go.

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    27/36

    CELLULAR REPLICASES 309

    ACKNOWLEDGMENTS

    The authors are grateful to the following individuals for helpful comments and

    discussions: Greg Bowman, Brian Chait, Megan Davey, David Jeruzalmi, John

    Kuriyan, and Alan Tackett. We are also grateful to Nina Yao for providing helpwith artwork. This work was supported by a grant from the National Institutes of

    Health (GM38839) and by the Howard Hughes Medical Institute.

    The Annual Review of Biochemistry is online at

    http://biochem.annualreviews.org

    LITERATURE CITED

    1. Kornberg A, Baker T. 1992. DNA Repli-

    cation. New York: Freeman. 931 pp. 2nd

    ed.

    2. Krishna TS, Kong XP, Gary S, Burg-

    ers PM, Kuriyan J. 1994. Cell 79:1233

    43

    3. Kong XP, Onrust R, ODonnell M,

    Kuriyan J. 1992. Cell 69:42537

    4. ODonnell M, Jeruzalmi D, Kuriyan J.

    2001. Curr. Biol. 11:R93546

    5. Marians KJ. 1992. Annu. Rev. Biochem.

    61:673719

    6. Kelman Z, ODonnell M. 1995. Annu.

    Rev. Biochem. 64:171200

    7. Benkovic SJ, Valentine AM, Salinas F.

    2001. Annu. Rev. Biochem. 70:181208

    8. McHenry CS. 2003. Mol. Microbiol. 49:

    115765

    9. Waga S, Stillman B. 1998. Annu. Rev.

    Biochem. 67:7215110. Burgers PM. 1998. Chromosoma 107:

    21827

    11. Hubscher U, Maga G, Spadari S. 2002.

    Annu. Rev. Biochem. 71:13363

    12. Forsburg SL. 2004. Microbiol. Mol. Biol.

    Rev. 68:10931

    13. Lehman IR, Kaguni LS. 1989. J. Biol.

    Chem. 264:426568

    14. Wold MS. 1997. Annu. Rev. Biochem. 66:

    619215. Takayama Y, Kamimura Y, Okawa M,

    Muramatsu S, Sugino A, Araki H. 2003.

    Genes Dev. 17:115365

    16. Kamimura Y, Masumoto H, Sugino A,

    Araki H. 1998. Mol. Cell. Biol. 18:6102

    9

    17. Kamimura Y, Tak YS, Sugino A, Araki

    H. 2001. EMBO J. 20:2097107

    18. Hoege C, Pfander B, Moldovan GL, Py-

    rowolakis G, Jentsch S. 2002. Nature

    419:13541

    19. Haracska L, Torres-Ramos CA, Johnson

    RE, Prakash S, Prakash L. 2004. Mol.

    Cell. Biol. 24:426774

    20. Kannouche PL, Wing J, Lehmann AR.

    2004. Mol. Cell 14:491500

    21. Warbrick E. 2000. BioEssays 22:997

    1006

    22. Lopez de Saro FJ, ODonnell M. 2001.

    Proc. Natl. Acad. Sci. USA 98:837680

    23. Maga G, Hubscher U. 2003. J. Cell Sci.

    116:305160

    24. Lopez de Saro FJ, Georgescu RE, Good-

    man MF, ODonnell M. 2003. EMBO J.22:640818

    25. Lindsey-Boltz LA, Bermudez VP, Hur-

    witz J, Sancar A. 2001. Proc. Natl. Acad.

    Sci. USA 98:1123641

    26. Naiki T, Kondo T, Nakada D, Matsumoto

    K, Sugimoto K. 2001. Mol. Cell. Biol. 21:

    583845

    27. Bellaoui M, Chang M, Ou J, Xu H, Boone

    C, Brown GW. 2003. EMBO J. 22:4304

    1328. Grabowski B, Kelman Z. 2003.Annu. Rev.

    Microbiol. 57:487516

    29. Tabor S, Huber HE, Richardson CC.1987.

    J. Biol. Chem. 262:1621223

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    28/36

    310 JOHNSON ODONNELL

    30. Huber HE,Tabor S, Richardson CC. 1987.

    J. Biol. Chem. 262:1622432

    31. Klemperer N, McDonald W, Boyle K,

    Unger B, Traktman P. 2001. J. Virol. 75:

    12298307

    32. McDonald WF, Klemperer N, Traktman

    P. 1997. Virology 234:16875

    33. Hernandez TR, Lehman IR. 1990. J. Biol.

    Chem. 265:1122732

    34. Gottlieb J, Marcy AI, Coen DM, Chall-

    berg MD. 1990. J. Virol. 64:597687

    35. Kaguni LS. 2004. Annu. Rev. Biochem.

    73:293320

    36. Kelly TJ, Brown GW. 2000. Annu. Rev.Biochem. 69:82980

    37. Bell SP, Dutta A. 2002. Annu. Rev.

    Biochem. 71:33374

    38. Kornberg T, Gefter ML. 1971. Proc. Natl.

    Acad. Sci. USA 68:76164

    39. McHenry CS, Crow W. 1979. J. Biol.

    Chem. 254:174853

    40. Studwell-Vaughan PS, ODonnell M.

    1993. J. Biol. Chem. 268:1178591

    41. Maki H, Kornberg A. 1985.J. Biol. Chem.260:1298792

    42. Studwell PS, ODonnell M. 1990. J. Biol.

    Chem. 265:117178

    43. Taft-Benz SA, Schaaper RM. 2004. J.

    Bacteriol. 186:277480

    44. Slater SC, Lifsics MR, ODonnell M,

    Maurer R. 1994. J. Bacteriol. 176:815

    21

    45. Braithwaite DK,Ito J. 1993.Nucleic Acids

    Res. 21:78780246. Steitz TA. 1999. J. Biol. Chem. 274:

    1739598

    47. Kim DR, McHenry CS. 1996. J. Biol.

    Chem. 271:20699704

    48. Kim DR, McHenry CS. 1996. J. Biol.

    Chem. 271:2069098

    49. Hamdan S, Carr PD, Brown SE, Ollis DL,

    Dixon NE. 2002. Structure 10:53546

    50. Keniry MA, Berthon HA, Yang JY, Miles

    CS, Dixon NE. 2000. Protein Sci. 9:72133

    51. DeRose EF, Darden T, Harvey S, Gabel

    S, Perrino FW, et al. 2003. Biochemistry

    42:363544

    52. Perrino FW, Harvey S, McNeill SM.

    1999. Biochemistry 38:160019

    53. Brenowitz S, Kwack S, Goodman MF,

    ODonnell M, Echols H. 1991. J. Biol.

    Chem. 266:788892

    54. Miller H, Perrino FW. 1996.Biochemistry

    35:1291925

    55. Fay PJ, Johanson KO, McHenry CS,Bam-

    bara RA. 1981. J. Biol. Chem. 256:976

    83

    56. LaDuca RJ, Fay PJ, Chuang C, McHenry

    CS, Bambara RA. 1983.Biochemistry 22:

    517788

    57. Stukenberg PT, Studwell-Vaughan PS,ODonnell M. 1991. J. Biol. Chem. 266:

    1132834

    58. Kuwabara N, Uchida H. 1981. Proc. Natl.

    Acad. Sci. USA 78:576467

    59. LaDuca RJ, Crute JJ, McHenry CS, Bam-

    bara RA. 1986. J. Biol. Chem. 261:7550

    57

    60. Reems JA, Wood S, McHenry CS. 1995.

    J. Biol. Chem. 270:560613

    61. Yao N, Leu FP, Anjelkovic J, TurnerJ, ODonnell M. 2000. J. Biol. Chem.

    275:1144050

    62. Naktinis V, Onrust R, Fang L, ODonnell

    M. 1995. J. Biol. Chem. 270:1335865

    63. Naktinis V, Turner J, ODonnell M. 1996.

    Cell 84:13745

    64. Leu FP, Hingorani MM, Turner J,

    ODonnell M. 2000. J. Biol. Chem. 275:

    3460918

    65. Stewart J, Hingorani MM, Kelman Z,ODonnell M. 2001. J. Biol. Chem. 276:

    1918289

    66. Turner J, Hingorani MM, Kelman Z,

    ODonnell M. 1999. EMBO J. 18:771

    83

    67. Jeruzalmi D, Yurieva O, Zhao Y, Young

    M, Stewart J, et al. 2001. Cell 106:41728

    68. Indiani C, ODonnell M. 2003. J. Biol.

    Chem. 278:4027281

    69. Pritchard AE, Dallmann HG, Glover BP,McHenry CS. 2000. EMBO J. 19:6536

    45

    70. Jeruzalmi D, ODonnell M, Kuriyan J.

    2001. Cell 106:42941

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    29/36

    CELLULAR REPLICASES 311

    71. Hingorani MM, ODonnell M. 1998. J.

    Biol. Chem. 273:2455063

    72. Tsuchihashi Z, Kornberg A. 1989.J. Biol.

    Chem. 264:1779095

    73. Xiao H, Naktinis V, ODonnell M. 1995.

    J. Biol. Chem. 270:1337883

    74. Guenther B, Onrust R, Sali A, ODonnell

    M, Kuriyan J. 1997. Cell 91:33545

    75. Onrust R, ODonnell M. 1993. J. Biol.

    Chem. 268:1176672

    76. Xiao H, Dong Z, ODonnell M. 1993. J.

    Biol. Chem. 268:1177984

    77. Kelman Z, Yuzhakov A, Andjelkovic J,

    ODonnell M. 1998. EMBO J. 17:243649

    78. Glover BP, McHenry CS. 1998. J. Biol.

    Chem. 273:2347684

    79. Olson MW, Dallmann HG, McHenry CS.

    1995. J. Biol. Chem. 270:2957077

    80. Ason B, Handayani R, Williams CR,

    Bertram JG, Hingorani MM, et al. 2003.

    J. Biol. Chem. 278:1003340

    81. Bloom LB, Turner J, Kelman Z, Beechem

    JM, ODonnell M, Goodman MF. 1996.J. Biol. Chem. 271:30699708

    82. Ason B, Bertram JG, Hingorani MM,

    Beechem JM, ODonnell M, et al. 2000.

    J. Biol. Chem. 275:300615

    83. Bertram JG, Bloom LB, Hingorani MM,

    Beechem JM, ODonnell M, Goodman

    MF. 2000. J. Biol. Chem. 275:28413

    20

    84. Williams CR, Snyder AK, Kuzmic P,

    ODonnell M, Bloom LB. 2004. J. Biol.Chem. 279:437685

    85. Bowman GD, ODonnell M, Kuriyan J.

    2004. Nature 429:72430

    86. Neuwald AF, Aravind L, Spouge JL,

    Koonin EV. 1999. Genome Res. 9:2743

    87. Davey MJ, Jeruzalmi D, Kuriyan J,

    ODonnell M. 2002. Nat. Rev. Mol. Cell

    Biol. 3:82635

    88. Dong Z, Onrust R, Skangalis M,

    ODonnell M. 1993. J. Biol. Chem. 268:1175865

    89. Johnson A, ODonnell M. 2003. J. Biol.

    Chem. 278:1440613

    90. Snyder AK, Williams CR, Johnson A,

    ODonnell M, Bloom LB. 2004. J. Biol.

    Chem. 279:438693

    91. Song MS, McHenry CS. 2001. J. Biol.

    Chem. 276:4870915

    92. Podobnik M, Weitze TF, ODonnell M,

    Kuriyan J. 2003. Structure 11:25363

    93. Goedken ER, Levitus M, Johnson A,

    Bustamante C, ODonnell M, Kuriyan J.

    2004. J. Mol. Biol. 336:104759

    94. Leu FP, ODonnell M. 2001. J. Biol.

    Chem. 276:4718594

    95. Wu YH, Franden MA, Hawker JR Jr,

    McHenry CS. 1984. J. Biol. Chem. 259:

    121172296. Flower AM, McHenry CS. 1986. Nucleic

    Acids Res. 14:8091101

    97. Tsuchihashi Z, Kornberg A. 1990. Proc.

    Natl. Acad. Sci. USA 87:251620

    98. Gao D, McHenry CS. 2001.J. Biol. Chem.

    276:444146

    99. Gao D, McHenry CS. 2001.J. Biol. Chem.

    276:443340

    100. Blinkova A, Hervas C, Stukenberg PT,

    Onrust R, ODonnell ME, Walker JR.1993. J. Bacteriol. 175:601827

    101. Onrust R, Finkelstein J, Turner J, Nakti-

    nis V, ODonnell M. 1995. J. Biol. Chem.

    270:1336677

    102. Glover BP, McHenry CS. 2000. J. Biol.

    Chem. 275:301720

    103. Yuzhakov A, Turner J, ODonnell M.

    1996. Cell 86:87786

    104. McHenry CS, Johanson KO. 1984. Adv.

    Exp. Med. Biol. 179:31519105. Johanson KO, McHenry CS. 1984.J. Biol.

    Chem. 259:458995

    106. McHenry CS. 1988. Biochim. Biophys.

    Acta 951:24048

    107. Wu CA, Zechner EL, Hughes AJ Jr, Fran-

    den MA, McHenry CS, Marians KJ. 1992.

    J. Biol. Chem. 267:406473

    108. Dallmann HG, Thimmig RL, McHenry

    CS. 1995. J. Biol. Chem. 270:2955562

    109. Glover BP, McHenry CS. 2001. Cell105:92534

    110. Kim S, Dallmann HG, McHenry CS, Mar-

    ians KJ. 1996. J. Biol. Chem. 271:21406

    12

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    30/36

    312 JOHNSON ODONNELL

    111. Kim S, Dallmann HG, McHenry

    CS, Marians KJ. 1996. Cell 84:643

    50

    112. Dallmann HG, Kim S, Pritchard AE, Mar-

    ians KJ, McHenry CS. 2000. J. Biol.

    Chem. 275:1551219

    113. Kowalczykowski SC. 2000. Trends Bio-

    chem. Sci. 25:15665

    114. Marians KJ. 2004. Philos. Trans. R. Soc.

    London Ser. B 359:7177

    115. Grompone G, Seigneur M, Ehrlich SD,

    Michel B. 2002. Mol. Microbiol. 44:

    133139

    116. Zechner EL, Wu CA, Marians KJ. 1992.J. Biol. Chem. 267:404553

    117. Zechner EL, Wu CA, Marians KJ. 1992.

    J. Biol. Chem. 267:405463

    118. Tougu K, Marians KJ. 1996. J. Biol.

    Chem. 271:21398405

    119. Tougu K, Marians KJ. 1996. J. Biol.

    Chem. 271:2139197

    120. Wu CA,Zechner EL, Reems JA, McHenry

    CS, Marians KJ. 1992. J. Biol. Chem.

    267:407483121. Stayton MM, Kornberg A. 1983. J. Biol.

    Chem. 258:1320512

    122. Sun W, Godson GN. 1996. J. Bacteriol.

    178:67015

    123. Sun W, Godson GN. 1998. J. Mol. Biol.

    276:689703

    124. Yuzhakov A, Kelman Z, ODonnell M.

    1999. Cell 96:15363

    125. Sinha NK, Morris CF, Alberts BM. 1980.

    J. Biol. Chem. 255:429093126. Chastain PD 2nd, Makhov AM, Nos-

    sal NG, Griffith J. 2003. J. Biol. Chem.

    278:2127685

    127. ODonnell ME. 1987. J. Biol. Chem.

    262:1655865

    128. Stukenberg PT, Turner J, ODonnell M.

    1994. Cell 78:87787

    129. Li X, Marians KJ. 2000. J. Biol. Chem.

    275:3475765

    130. Leu FP, Georgescu R, ODonnell M.2003. Mol. Cell 11:31527

    131. Wu CA, Zechner EL, Marians KJ. 1992.

    J. Biol. Chem. 267:403044

    132. Higuchi K, Katayama T, Iwai S, Hidaka

    M,HoriuchiT,MakiH.2003. Genes Cells

    8:43749

    133. McInerney P, ODonnell M. 2004.J. Biol.

    Chem. 279:2154351

    134. Lopez de Saro FJ, Georgescu RE,

    ODonnell M. 2003. Proc. Natl. Acad. Sci.

    USA 100:1468994

    135. Bunting KA, Roe SM, Pearl LH. 2003.

    EMBO J. 22:588392

    136. Burnouf DY, Olieric V, Wagner J, Fujii

    S, Reinbolt J, et al. 2004. J. Mol. Biol.

    335:118797

    137. Duzen JM, Walker GC, Sutton MD. 2004.

    DNA Repair3:30112138. Miyachi K, Fritzler MJ, Tan EM. 1978. J.

    Immunol. 121:222834

    139. Tan CK, Castillo C, So AG, Downey

    KM. 1986. J. Biol. Chem. 261:12310

    16

    140. Prelich G, Kostura M, Marshak DR,

    Mathews MB, Stillman B. 1987. Nature

    326:47175

    141. Prelich G, Tan CK, Kostura M, Mathews

    MB, So AG, et al. 1987. Nature 326:51720

    142. Tsurimoto T, Stillman B. 1990. Proc.

    Natl. Acad. Sci. USA 87:102327

    143. Yao N, Turner J, Kelman Z, Stukenberg

    PT, Dean F, et al. 1996. Genes Cells 1:

    10113

    144. Mossi R, Jonsson ZO, Allen BL, Hardin

    SH, Hubscher U. 1997. J. Biol. Chem.

    272:176976

    145. Oku T, Ikeda S, Sasaki H, Fukuda K,Morioka H, et al. 1998. Genes Cells 3:

    35769

    146. Gulbis JM, Kelman Z, Hurwitz J,

    ODonnell M, Kuriyan J. 1996. Cell 87:

    297306

    147. Shamoo Y, Steitz TA. 1999. Cell 99:155

    66

    148. Franklin MC, Wang J, Steitz TA. 2001.

    Cell 105:65767

    149. Hingorani MM, ODonnell M. 2000.Curr. Biol. 10:R2529

    150. Dalrymple BP, Kongsuwan K, Wijffels G,

    Dixon NE, Jennings PA. 2001. Proc. Natl.

    Acad. Sci. USA 98:1162732

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    31/36

    CELLULAR REPLICASES 313

    151. Fairman M, Prelich G, Tsurimoto T, Still-

    man B. 1988. Biochim. Biophys. Acta

    951:38287

    152. Waga S, Stillman B. 1994. Nature

    369:20712

    153. Lee SH, Eki T, Hurwitz J. 1989. Proc.

    Natl. Acad. Sci. USA 86:736165

    154. Mossi R, Hubscher U. 1998. Eur. J.

    Biochem. 254:20916

    155. Majka J, Burgers PM. 2004. Prog. Nucleic

    Acid Res. Mol. Biol. 78:22760

    156. ODonnell M, Onrust R, Dean FB, Chen

    M, Hurwitz J. 1993. Nucleic Acids Res.

    21:13157. Cullmann G, Fien K, Kobayashi R, Still-

    man B. 1995.Mol. Cell. Biol. 15:466171

    158. Tsurimoto T, Stillman B. 1991. J. Biol.

    Chem. 266:195060

    159. Podust VN, Tiwari N, Stephan S, Fanning

    E. 1998. J. Biol. Chem. 273:3199299

    160. Uhlmann F, Gibbs E, Cai J, ODonnell

    M, Hurwitz J. 1997. J. Biol. Chem.

    272:1006571

    161. Bunz F, Kobayashi R, Stillman B. 1993.Proc. Natl. Acad. Sci. USA 90:1101418

    162. Gomes XV, Gary SL, Burgers PM. 2000.

    J. Biol. Chem. 275:1454149

    163. Uhlmann F, Cai J, Gibbs E, ODonnell

    M, Hurwitz J. 1997. J. Biol. Chem.

    272:1005864

    164. Yao N, Coryell L, Zhang D, Georgescu

    RE, Finkelstein J, et al. 2003. J. Biol.

    Chem. 278:5074453

    165. Fotedar R, Mossi R, Fitzgerald P, Rous-selle T, Maga G, et al. 1996. EMBO J.

    15:442333

    166. Cai J, Gibbs E, Uhlmann F, Phillips B, Yao

    N, et al. 1997. J. Biol. Chem. 272:18974

    81

    167. Gomes XV, Schmidt SL, Burgers PM.

    2001. J. Biol. Chem. 276:3477683

    168. Podust VN, Tiwari N, Ott R, Fanning E.

    1998. J. Biol. Chem. 273:1293542

    169. Cai J, Yao N, Gibbs E, Finkelstein J,Phillips B, et al. 1998. Proc. Natl. Acad.

    Sci. USA 95:1160712

    170. Schmidt SL, Gomes XV, Burgers PM.

    2001. J. Biol. Chem. 276:3478491

    171. Gomes XV, Burgers PM. 2001. J. Biol.

    Chem. 276:3476875

    172. Hingorani MM, Coman MM. 2002. J.

    Biol. Chem. 277:4721324

    173. Conaway RC, Lehman IR. 1982. Proc.

    Natl. Acad. Sci. USA 79:252327

    174. Conaway RC, Lehman IR. 1982. Proc.

    Natl. Acad. Sci. USA 79:458588

    175. Tsurimoto T, Stillman B. 1991. J. Biol.

    Chem. 266:196168

    176. Maga G, Hubscher U. 1996.Biochemistry

    35:576477

    177. Yuzhakov A, Kelman Z, Hurwitz J,

    ODonnell M. 1999. EMBO J. 18:618999

    178. Zuo S, Bermudez V, Zhang G, Kelman Z,

    HurwitzJ.2000.J. Biol. Chem. 275:5153

    62

    179. Podust VN, Chang LS, Ott R, Dianov

    GL, Fanning E. 2002. J. Biol. Chem.

    277:3894901

    180. Burgers PM, Gerik KJ. 1998. J. Biol.

    Chem. 273:1975662

    181. Gerik KJ, Li X, Pautz A, BurgersPM. 1998. J. Biol. Chem. 273:19747

    55

    182. MacNeill SA, Baldacci G, Burgers PM,

    Hubscher U. 2001. Trends Biochem. Sci.

    26:1617

    183. Bermudez VP, MacNeill SA, Tappin I,

    Hurwitz J. 2002. J. Biol. Chem. 277:

    3685362

    184. Johansson E, Majka J, Burgers PM. 2001.

    J. Biol. Chem. 276:4382428185. Lee MY, Tan CK, Downey KM, So AG.

    1984. Biochemistry 23:190613

    186. MacNeill SA, Moreno S, Reynolds N,

    Nurse P, Fantes PA. 1996. EMBO J.

    15:461328

    187. Sanchez Garcia J, Ciufo LF, Yang X,

    Kearsey SE, MacNeill SA. 2004. Nucleic

    Acids Res. 32:300516

    188. Chen X, Zuo S, Kelman Z, ODonnell M,

    Hurwitz J, Goodman MF. 2000. J. Biol.Chem. 275:1767782

    189. Mozzherin DJ, McConnell M, Jasko MV,

    Krayevsky AA, Tan CK, et al. 1996. J.

    Biol. Chem. 271:3171117

    byUniversidadeFederaldoRiodeJaneiroon08/16/11.Forpersonaluseonly.

  • 8/4/2019 Annual Reviews - Replication

    32/36

    314 JOHNSON ODONNELL

    190. Hashimoto K, Shimizu K, Nakashima N,

    Sugino A. 2003. Biochemistry 42:14207

    13

    191. Mozzherin DJ, Tan CK, Downey KM,

    Fisher PA. 1999. J. Biol. Chem. 274:

    1986267

    192. Einolf HJ, Guengerich FP. 2000. J. Biol.

    Chem. 275:1631622

    193. Reynolds N, Warbrick E, Fantes PA, Mac-

    Neill SA. 2000. EMBO J. 19:110818

    194. Johansson E, Garg P, Burgers PM. 2004.

    J. Biol. Chem. 279:190715

    195. Morrison A, Araki H, Clark AB,

    Hamatake RK, Sugino A. 1990. Cell 62:114351

    196. Feng W, Rodriguez-Menocal L, Tolun G,

    DUrso G. 2003.Mol. Biol. Cell 14:3427

    36

    197. Aparicio OM, Weinstein DM, Bell SP.

    1997. Cell 91:5969

    198. Navas TA, Zhou Z, Elledge SJ. 1995. Cell

    80:2939

    199. Navas TA, Sanchez Y, Elledge SJ. 1996.

    Genes Dev. 10:263243200. DUrso G, Nurse P. 1997. Proc. Natl.

    Acad. Sci. USA 94:1249196

    201. Kesti T, Flick K, Keranen S, Syvaoja JE,

    W