DataPDF Available

Archaeal DNA Replication_120 Annu Rev Genet 2014 Kelman & Kelman

Authors:
GE48CH04-Kelman ARI 13 October 2014 8:44
Archaeal DNA Replication
Lori M. Kelman1and Zvi Kelman2
1Program in Biotechnology, Montgomery College, Germantown, Maryland 20876;
email: lori.kelman@montgomerycollege.edu
2National Institute of Standards and Technology and Institute for Bioscience and Biotechnology
Research, Rockville, Maryland 20850; email: zkelman@umd.edu
Annu. Rev. Genet. 2014. 48:71–97
The Annual Review of Genetics is online at
genet.annualreviews.org
This article’s doi:
10.1146/annurev-genet-120213-092148
Copyright c
2014 by Annual Reviews.
All rights reserved
Keywords
Archaea, initiation of DNA replication, protein-DNA interactions,
protein-protein interactions, regulation of DNA replication, protein
structure
Abstract
DNA replication is essential for all life forms. Although the process is funda-
mentally conserved in the three domains of life, bioinformatic, biochemical,
structural, and genetic studies have demonstrated that the process and the
proteins involved in archaeal DNA replication are more similar to those in
eukaryal DNA replication than in bacterial DNA replication, but have some
archaeal-specific features. The archaeal replication system, however, is not
monolithic, and there are some differences in the replication process between
different species. In this review, the current knowledge of the mechanisms
governing DNA replication in Archaea is summarized. The general features
of the replication process as well as some of the differences are discussed.
71
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
Click here for quick links to
Annual Reviews content online,
including:
• Other articles in this volume
• Top cited articles
• Top downloaded articles
• Our comprehensive search
Further
ANNUAL
REVIEWS
GE48CH04-Kelman ARI 13 October 2014 8:44
INTRODUCTION
In 1977, Carl Woese and colleagues recognized that the rRNA sequences and other physiological
features of a group of prokaryotes were different enough from those in Bacteria to classify them as a
separate domain: the Archaebacteria (now Archaea) (177). The archaeal domain is currently divided
into six different phyla (or kingdoms): Aigarchaeota, Crenarchaeota, Euryarchaeota, Korarchaeota,
Nanoarchaeota, and Thaumarchaeota (50).
Early bioinformatic studies using complete archaeal genome sequences suggested that although
Archaea contain circular chromosomes, as do Bacteria, the proteins and complexes that participate
in DNA replication are more closely related to those of Eukarya than to those of Bacteria (Table 1).
Since then, phylogenetic, biochemical, structural, and genetic studies have demonstrated the rela-
tionship between the archaeal and eukaryal DNA replication systems (47, 50, 60, 61, 71, 127, 143).
Here, the current state of knowledge on the mechanism of DNA replication in Archaea, the
similarities and differences between species, and the properties of individual proteins and com-
plexes are discussed. Because of space limitations, only representative references are provided, and
when appropriate the reader is referred to other reviews.
DNA replication takes place during the S phase of the cell cycle and is tightly coordinated with
other cell-cycle events and cell division. This review concentrates on replication, and the reader
is referred to several reviews for information on other aspects of the archaeal cell cycle (103, 113,
153). Topoisomerases play a major role in DNA replication by untwisting the DNA in front of and
behind the moving replication fork. Topoisomerases are not covered here but have been reviewed
(21, 42).
OVERVIEW OF DNA REPLICATION
Chromosomal DNA replication is an essential process that ensures the accurate and timely dupli-
cation of the genetic information. It is a complex and highly regulated process that is functionally
and structurally conserved in all life forms (127). In all organisms, the process, which occurs during
S phase, begins at a specific region on the chromosome called the origin of replication. Origin
binding proteins (OBPs) bind to the origin and locally unwind it (35) and, together with the he-
licase loader, assemble the two replicative helicases around the DNA to form the prereplication
complex (pre-RC) (Figure 1). Additional proteins bind to the pre-RC to form the preinitiation
complex (pre-IC), in which the DNA duplex is unwound and single-stranded (ss) DNA binding
protein (SSB) coats the ssDNA. DNA primase, DNA polymerase, and the rest of the replication
machinery are recruited to the SSB-ssDNA nucleofilament at the origin. Upon activation, the
two helicases begin to unwind the duplex in opposite directions using energy derived from nu-
cleoside triphosphate hydrolysis. The moving helicases form the two replication forks to initiate
bidirectional DNA synthesis. The ssDNA exposed behind the helicase is coated with SSB. Ow-
ing to the antiparallel nature of DNA and the unidirectionality of DNA polymerase, the leading
strand is synthesized continuously, whereas the lagging strand is copied discontinuously as a se-
ries of Okazaki fragments (Figure 2,Figure 3a). A number of enzymes, including a nuclease,
a DNA polymerase, and DNA ligase, process the Okazaki fragments to form the mature duplex
DNA (Figure 3b). At the end of the replication process, the two replication forks (on a circular
chromosome, or adjacent forks on a linear chromosome) collide and DNA synthesis terminates.
ORIGIN OF REPLICATION
In all organisms, DNA replication starts at specific sites on the chromosome known as origins
of replication. All origins contain similar characteristics, including one or more A/T-rich regions
72 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
Table 1 A summary of DNA replication features in the three domains of lifea
Archaeab
Attribute Bacteria Eukarya Euryarchaea Crenarchaea
Chromosome Circular Linear Circular Circular
Replication origin Single Multiple Single or Multiple Single or Multiple
Prereplication complex (pre-RC)
Origin recognition DnaA (1) ORC (6) Cdc6c(1) Cdc6 (1)
WhiPd(1)
HelicaseeDnaB (1) MCM (6) MCM (1) MCM (1)
Helicase loadereDnaC (1) Cdc6 (1)
Cdt1 (1)
Cdc6 (1) Cdc6 (1)
Preinitiation complex (pre-IC)
Cdc45 -Cdc45 (1) Cdc45 (1) Cdc45 (1)
GINS -GINS (4) GINS (1–2) GINS (1–2)
Single-stranded DNA binding
protein (SSB)
SSB (1) RPA (3) RPA (1–3) SSB (1)f
Elongation complex
Primase DnaG (1) Polα/primase (4) Primase (2) Primase (2)
Sliding clamp β-subunit (1) PCNA (1) PCNA (1) PCNA (3)
Clamp loader τ-complex (5) RFC (5) RFC (2) RFC (2)
DNA polymerase PolC (3) PolBg(1) PolB/PolD (1h/2)PolB (2–3)
Okazaki fragment maturation
Primer removal PolI (1) Fen1 (1) Fen1 (1) Fen1 (1)
Gap filling PolI (1) PolBi(1) PolB/PolD (1h/2)PolB (2–3)
DNA ligase NAD+-dependent (1) ATP-dependent (1) ATP-dependentj(1) ATP-dependent (1)
aBacteria-like features and proteins are in red, eukaryal-like features and proteins are in blue, and archaeal-specific proteins are in green. The range of
homologs identified in different species is shown in parentheses.
bTwo representative kingdoms are shown.
cThe genomes of species belonging to Methanococcales and Methanopyrales do not contain genes encoding for Cdc6 homologs.
dAn archaeal homolog of the eukaryotic Cdt1 protein, WhiP, was identified in some species.
eIn bacteria, the helicase and helicase loader are not considered to be part of the pre-RC but rather the pre-IC. Because this paper is about archaea, these
proteins are included under pre-RC.
fThe crenarchaeal SSB shares some features with the bacterial SSB and other features with the eukaryotic replication protein A (RPA).
gAll three replicative DNA polymerases in eukarya (Polα,Polδ,andPolε) belong to family B.
hIn some archaeal species, PolB is not essential for cell viability.
iPolδ, the lagging-strand DNA polymerase, participates in Okazaki fragment maturation.
jSome archaeal DNA ligases use NAD+as a cofactor.
Abbreviations: MCM, minichromosome maintenance; ORC, origin recognition complex; PCNA, proliferating cell nuclear antigen; RFC, replication
factor C.
referred to as duplex unwinding elements (DUE), and specific sequences that facilitate the binding
of OBPs. Upon binding to the origin, the proteins melt the duplex DNA to form the initial
replication bubble (Figure 1) (15).
Origins of replication in Archaea were first identified using in silico skew analysis (74, 105). This
method made use of an observation in Bacteria that strand-specific biases in nucleotide, oligomer,
and codon frequencies could be identified along the chromosome, with an abrupt change in the
bias at the origin of replication and termination regions.
The origin of Pyrococcus abyssi was the first to be mapped in vivo using pulsed-field gel elec-
trophoresis and two-dimensional gel analysis (118, 126). As predicted by skew analysis (105),
www.annualreviews.org Archaeal DNA Replication 73
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
Origin
c
b
e
d
f
a
Cdc6
MCM
Primase
RNA primer
DNA polymerase
PCNA
Figure 1
Initiation of DNA replication. The (a) origin of replication is (b) recognized by Cdc6 ( yellow). (c) The binding
of Cdc6 to the origin region forms the initial replication bubble. (d) Cdc6 participates in the assembly of the
two m inichromosome maintenance (MCM) helicases ( green) around the DNA to form the two replication
forks. (e) The helicases move in opposite directions, enlarging the bubble. Primase ( pink) is recruited to the
DNA and synthesizes the RNA primers (light green). ( f) The rest of the replication machinery assembles and
initiates bidirectional DNA synthesis. For simplicity, the single-stranded DNA binding protein is omitted
and only the DNA polymerase (orange) and proliferating cell nuclear antigen (PCNA) (blue)areshown.
P. abyssi contains a single origin of replication (referred to as oriC) located in an intergenic region
of the chromosome. Bidirectional DNA synthesis initiates from the origin and terminates in a
region of the chromosome opposite the origin (119, 126).
The presence of a single origin in a circular archaeal genome was not surprising, because
bacterial circular chromosomes contain a single origin. However, skew analysis failed to identify a
clear origin of replication in some archaeal species, and the use of more refined in silico algorithms
suggested the presence of multiple origins in certain species (184). The presence of multiple
74 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
MCM
Lagging strand
Leading strand
Primase
Primer
DNA polymerase
GINS
RPA
RFC
PCNA
Figure 2
A schematic representation of the archaeal replisome. MCM (minichromosome maintenance) is green, DNA
polymerase is orange, primase is pink, RFC (replication factor C) is purple, GINS is gray, RPA (replication
protein A) is yellow, and PCNA (proliferating cell nuclear antigen) is blue. Primers on the lagging strand are
shown in light green.
origins could confound skew analysis and explain the failure to identify an origin. Studies using
two-dimensional gel analysis, marker frequency analysis, and replication initiation point (RIP)
mapping of two Sulfolobus species revealed that these organisms contain three origins of replication
(72, 106, 149). Subsequent studies showed that other archaeal species also contain multiple origins.
Archaeal origins have characteristics similar to origins of replication from other organisms
(15). They are located in intergenic regions, are rich in A and T nucleotides, contain one or
more DUEs, and contain binding sites for OBPs [referred to as origin recognition boxes (ORBs)].
Many origins are located in regions of the chromosome that encodes DNA replication proteins.
Many archaeal origins are located upstream from genes that encode the archaeal homolog of the
eukaryotic Cdc6 protein. The Cdc6 proteins are the archaeal OBPs and were shown to bind to
ORBs. The proximity of the origin of replication to the OBPs is also common in Bacteria (89).
Several hypotheses have been proposed to explain the proximity of the gene that encodes the OBP
to the origin, such as the ability of the proteins to associate with the origin as soon as they are
synthesized or to reduce the probability of loss of the gene that encodes the OBP due to genomic
rearrangement (71, 103).
In several archaeal species with multiple origins of replication, the origin is located in the
vicinity of the homolog of another eukaryotic replication protein, Cdt1. This protein binds to the
origin of replication in its vicinity (148, 154). Several origins are not located upstream of a known
initiator protein (138), and the proteins that bind these origins have not yet been identified.
The origins of replication from most Archaea studied contain clear ORBs. ORBs are inverted
repeats located on both sides of the DUE region(s) and were shown to be the binding site for
the Cdc6 proteins (36, 45, 174). The ORBs from different species share sequence similarity with
www.annualreviews.org Archaeal DNA Replication 75
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
I
II
III
IV
b
DNA polymerase
PCNA
Fen1
DNA ligase
RPA
a
I
II
III
IV
V
Primase
RNA primer
PCNA
RFC
DNA polymerase
RPA
R
F
Figure 3
Okazaki fragment synthesis and maturation. (a) Okazaki fragment synthesis. I. Primase ( pink) synthesizes an RNA primer (light green)
on RPA (replication protein A) ( yellow)-coated single-stranded DNA. II. This primer is recognized by RFC (replication factor C)
(purple). III. RFC assembles PCNA (proliferating cell nuclear antigen) (blue) around the primer. IV. The DNA polymerase (orange)
associates with PCNA. V. This association initiates processive DNA synthesis. (b) Okazaki fragment maturation. I. The lagging-
strand DNA polymerase (orange) elongates the Okazaki fragment with the assistance of PCNA (blue). II. When the DNA polymerase
reaches the previous Okazaki fragment, its strand displacement activity forms a flap structure. III. The flap is removed by the activity of
Fen1 (flap endonuclease 1) (teal) and PCNA. IV. The nick remaining on the DNA is sealed by DNA ligase ( yellow) and PCNA.
a consensus sequence referred to as the mini-ORB, followed, in some origins, by a string of
G nucleotides (a G-string) to form the intact ORB (10, 27). It was shown that the mini-ORB
is sufficient to bind Cdc6 proteins and that Cdc6 from one organism can bind an ORB from
another species (149). Some origins do not have a clear ORB or mini-ORB consensus sequence,
although they contain other inverted repeats (138). It is not clear whether Cdc6, Cdt1, or other
species-specific initiator proteins can bind those repeats.
Interestingly, in some species the origins of replication could be deleted with no loss of viability
(53). However, in the absence of a canonical origin, the cellular recombination machinery appears
to be required for cell viability. The mechanism of initiation in this case is not clear. It is also not
known how widespread this phenomenon is among the Archaea.
PREREPLICATION COMPLEX
The pre-RC is the protein complex that assembles at the origin of replication during the G1 phase
of the cell cycle and is responsible for the regulation of the initiation process. In Eukarya, the pre-
RC is composed of four main components: the six-subunit origin recognition complex (ORC),
Cdc6 and Cdt1 proteins, and the six-subunit minichromosome maintenance (MCM) helicase. In
Archaea, pre-RC formation includes the assembly of the Cdc6 protein and the MCM helicase. In
some Archaea, a Cdt1 homolog (WhiP) or other, not yet identified, proteins may also be part of
the pre-RC.
76 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
Cdc6
The OBP in Bacteria is the DnaA protein, which binds the DnaA box (the bacterial ORB) at the
origin of replication. In Eukarya, the OBP is the hexameric ORC complex composed of the Orc1–6
proteins. ORC associates with the origin and recruits additional components of the pre-RC, such
as the MCM helicase and the Cdc6 protein, to participate in MCM assembly at the origin. The
eukaryotic Orc1–5 proteins, Cdc6, and the bacterial DnaA protein all belong to the initiator clade
of the AAA+family of ATPases (27), and Orc1–5 and Cdc6 were proposed to be paralogs (110).
In nearly all Archaea, at least one clear homolog of the eukaryotic Orc or Cdc6 proteins has
been identified, and most contain two or three genes encoding for the protein (143). Owing to
similarity to both the eukaryotic Orc and Cdc6 proteins, the archaeal enzyme has been called Cdc6,
Orc, Cdc6/Orc, and Cdc6/Orc1. In this review, the protein is referred to as Cdc6. Although the
Cdc6 protein is essential for cell viability in most species studied (154), it appears to be dispensable
in others (T. Santangelo & Z. Kelman, unpublished results).
Because of their similarity to both the ORC complex and the Cdc6 protein, the archaeal
enzymes were suggested to be involved in both origin recognition and helicase loading (Table 1;
Figure 1). The role of Cdc6 protein in origin recognition has been established. The first suggestion
of the association of Cdc6 with the origin of replication came from chromatin immunoprecipitation
(ChIP) analysis with P. abyssi (118). Subsequent in vivo and in vitro studies illustrated that although
the proteins can bind both ssDNA and double-stranded (ds) DNA, they show clear preference for
the ORB dsDNA (10, 27).
Although the archaeal Cdc6 proteins bind origins, the mechanism by which they initiate DNA
replication is not clear. The bacterial DnaA protein unwinds origin duplex DNA to facilitate the
assembly of the replicative helicase DnaB on the exposed ssDNA (27). In contrast, binding of the
eukaryotic ORC to the origin does not induce unwinding. Furthermore, there is accumulating
evidence to suggest that the eukaryotic MCM assembles at the origin around dsDNA and not
around ssDNA, as suggested for Bacteria (27). The similarities between the archaeal Cdc6 and the
eukaryotic ORC and Cdc6 proteins and the lack of clear data on the ability of the archaeal Cdc6
protein to melt the origin may suggest that, as in Eukarya, during pre-RC formation the MCM
helicase is assembled on dsDNA and origin melting occurs during pre-IC formation.
Eukaryotic Cdc6 plays an essential role in the assembly of the MCM helicase at the origin,
and thus it was suggested that the archaeal Cdc6 plays a similar role. Direct interaction between
Cdc6 and MCM has been reported in several archaeal species, and the interaction was capable of
regulating helicase activity (152). Studies showed that the interaction between Cdc6 and MCM
inhibited helicase activity (158), which is likely due to the dissociation of the MCM hexamers in
the presence of Cdc6 (79, 159). The inhibitory effect of Cdc6 on helicase activity is reminiscent
of the observation made in bacteria, where binding of the helicase loader DnaC to the helicase
DnaB inhibits helicase activity (172). Thus, this observation may support the hypothesis that the
archaeal Cdc6 functions as a helicase loader. Thermoplasma acidophilum Cdc6, however, stimulates
the activity of the MCM helicase (51, 52), which may suggest a different mechanism of helicase
loading in that species. Nevertheless, the mechanism of helicase assembly at the archaeal origin
and the role of Cdc6 protein in the process are not known, although several mechanisms have
been proposed (151).
Cdc6, DnaA, and Orc1-5 belong to the AAA+superfamily of ATPases (27, 34). The structural
organization of the Cdc6 protein is similar to that of the bacterial DnaA proteins in which the
N-terminal AAA+catalytic unit is followed by a DNA binding domain. However, although in
DnaA the DNA binding domain is a helix-turn-helix (HTH), the Cdc6 protein contains a winged-
helix domain (WHD) (27). Biochemical studies demonstrated that an intact WHD is required for
www.annualreviews.org Archaeal DNA Replication 77
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
ORB binding (10, 27, 35). The WHD also participates in the interaction between the Cdc6 and
MCM proteins (30, 66).
In addition to all known motifs present in other members of the AAA+family of ATPases,
an α-helix inserted into the canonical AAA+fold was also revealed by the structure of Cdc6 and
DnaA. This helical insertion is found in all proteins belonging to the initiator clade of the AAA+
family and is referred to as the initiator specific motif (ISM) (27).
The three-dimensional structure of Cdc6 bound to the ORB has been determined (36, 45). In
addition to extensive contact between the WHD and the DNA, the ISM also contacts the DNA.
However, although ORB sequences are conserved and in vivo and in vitro studies showed the
specificity of Cdc6 binding to the ORBs, only a few sequence-specific contacts were observed in
the co-crystals (36, 45). The mechanism by which binding specificity is achieved is not yet known
and may involve the helicase or other proteins (10, 27, 174).
Cdc6 is expressed during the G1 phase and degraded after initiation (107, 149). However,
the regulation of expression and of subsequent degradation is not understood. In Eukarya, phos-
phorylation of Cdc6 following the initiation process marks the enzyme for degradation. A similar
mechanism may regulate Cdc6 in Archaea. A number of archaeal protein kinases are cell-cycle
regulated (107), and in vivo phosphorylation of Cdc6 protein has been reported (144).
Alternatively, the archaeal Cdc6 proteins can autophosphorylate in vitro using the γ-phosphate
of ATP (46). Studies suggest that autophosphorylation may be regulated by MCM binding to Cdc6
and by the association of Cdc6 with DNA (46, 66). Although the role of autophosphorylation in
Cdc6 function is not known, it may play a regulatory role during pre-RC formation (47). It will
be important to determine whether Cdc6 can autophosphorylate in vivo as a first step to establish
whether the process has a physiological role.
In bacteria, the activity of DnaA is regulated by a mechanism referred to as regulatory inactiva-
tion of DnaA (RIDA), which ensures that the chromosome replicates only once per cell cycle (98).
RIDA is dependent upon interaction between the bacterial processivity factor of DNA polymerase
III (PolIII) holoenzyme (the β-subunit), DnaA, and a homologous-to-DnaA (Hda) protein. Inter-
action between Hda and DnaA stimulates ATP hydrolysis by DnaA, resulting in an ADP-bound
DnaA protein that cannot bind to the origin of replication. However, Hda must associate with
the β-subunit on DNA to stimulate ATP hydrolysis by DnaA. Proliferating cell nuclear antigen
(PCNA) is the archaeal functional homolog of the β-subunit (80). In several species, the Cdc6
protein contains a PCNA-interacting protein (PIP) motif required for interactions with PCNA,
and PCNA and Cdc6 can form a complex in vivo (102) and in vitro (3). It was therefore proposed
that regulation of Cdc6 function in Archaea might resemble the bacterial RIDA process (100).
However, no homolog of Hda has been identified in Archaea. It is possible that another unidenti-
fied protein is the archaeal functional homolog of Hda, or that an Hda-like protein is not needed
in Archaea as Cdc6 can interact directly with PCNA.
Other mechanisms may also regulate Cdc6 function. Archaeal Cdc6 proteins crystallized as
monomers with ADP tightly bound at the active site, and denaturation of Cdc6 was required for
removal of the bound ADP. As it is thought that the Cdc6 protein is active in the ATP-bound
form, several hypotheses for the role of the tight ADP binding for regulating the enzyme have
been proposed (10).
Winged-Helix Initiator Protein
Cdt1 is a eukaryotic protein that participates in the initiation of DNA replication as a part of the
pre-RC (27). Structurally, the middle and the C-terminal portions of the eukaryotic Cdt1 protein
have a WHD (97). When the three origins of replication in Sulfolobus solfataricus were identified, it
78 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
was noted that only two origins are located upstream to a gene encoding Cdc6 (106). However, the
protein encoded by the gene upstream of the third S. solfataricus origin also contains two WHD,
although at the N and C termini of the molecule. This suggested that the protein might be the
archaeal homolog of Cdt1 (148). Because amino acid similarity between the archaeal and eukaryal
enzymes is low, the archaeal protein is called WhiP (winged-helix initiator protein). It was found
that the WhiP protein could bind to the origin of replication near its gene (148). To date, WhiP
homologs have been identified only in a small subset of species belonging to the crenarchaeota
branch. However, as many archaeal genomes contain proteins with a single WHD that is similar
to that found in WhiP, these proteins may play a role in the initiation process in those species.
Minichromosome Maintenance Helicase
The MCM proteins are the replicative helicases in Archaea responsible for the separation of the
dsDNA in front of the replication fork. The archaeal MCM proteins have been extensively studied
using structural, biochemical, and genetic tools (11, 16, 28, 93, 123, 152, 164).
In Eukarya, MCM is a complex of six different but related proteins (Mcm2–7). Most archaeal
species contain a single gene that encodes MCM. To date, in archaeal species with multiple MCM
genes, only one is essential for viability (59, 133). The single MCM homolog from the archaeon
Methanothermobacter thermautotrophicus was the first to be biochemically characterized (23, 79,
156), but others have since been studied. As with other helicases, the MCM utilizes energy from
ATP hydrolysis to translocate along one strand of the DNA and displace the other. The enzyme
binds to and translocates along ssDNA and dsDNA in the 3to 5direction, and in vitro the enzyme
is processive on its own and able to displace several hundred bases (16, 28, 93, 152, 164). However,
on dsDNA the enzyme probably interacts with one strand, resulting in the observed 3to 5direc-
tionality. Biochemical and biophysical studies with mutated enzymes suggested that ssDNA and
dsDNA are located in the central hole of the hexameric ring. During DNA unwinding, the helicase
may encounter proteins on DNA (such as histones) or short RNA molecules, forming R-loops.
Archaeal MCM can displace proteins from DNA and unwind DNA-RNA hybrids (161, 162).
Early studies suggested that archaeal MCM proteins form dodecameric structures in solution
as judged by size-exclusion chromatography and sedimentation methods (23, 79, 156). Further
support for dodecameric structures came from the three-dimensional structure of the N-terminal
portion of the enzyme (40, 104) and the full-length protein (91). Electron microscope reconstruc-
tion studies, however, indicated that the enzyme could form other structures, including hexamers,
heptamers, octamers, filaments, and open rings (28, 152, 164).
Although the ability of MCM to form both hexamer and head-to-head dodecameric structures
was first reported with an archaeal enzyme, it was later shown for the eukaryotic MCM (145) and
the bacterial DnaB (166). It is thought that the replicative helicases are assembled as dodecameric
rings at the origin. It is not clear whether the two hexamers associate during elongation or are
separated during DNA synthesis and move away from each other (151, 168). However, in vitro
studies indicate that the archaeal enzymes are active as single hexamers (160), suggesting that upon
loading the two rings work independently.
Structurally, archaeal MCM proteins can be divided into three main parts: an N-terminal part,
a catalytic region, and a C-terminal HTH domain (40, 152, 164). The N-terminal part of MCM
forms hexamers and dodecamers on its own (40, 67, 92). This part of the molecule contains a C4-
type zinc finger and a β-hairpin motif shown to be involved in MCM interaction with DNA (78,
141). The structure also revealed the presence of a long loop shown to play a role in communication
between the N-terminal DNA binding region and the catalytic part of the molecule (6, 150). The
central part of MCM contains the catalytic domains and all the conserved motifs found in other
www.annualreviews.org Archaeal DNA Replication 79
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
AAA+family members (28, 152, 164). The structure also revealed the presence of three β-hairpins
in the catalytic part; all are needed for helicase activity (17, 63, 122). The C-terminal part of MCM
was suggested to fold into an HTH domain (2), although the domain was not seen clearly in any
of the structures of the molecule. Biochemical studies suggested that the domain might play a
regulatory role, as removal resulted in increased helicase activity in vitro (7).
PREINITIATION COMPLEX
During S phase of the cell cycle, the pre-RC recruits additional factors, including Cdc45 and
GINS, to the origin to establish the pre-IC. Following pre-IC formation, the dsDNA is unwound
and SSB associates with the exposed ssDNA. The replisome is assembled at the origin and DNA
replication initiates.
Single-Stranded DNA Binding Protein
SSB plays an essential role in DNA replication by coating ssDNA behind the helicase to prevent
reannealing and protect the ssDNA from attack by nucleases and chemical modification (32).
In Bacteria and Eukarya, SSB also stimulates the activity of DNA polymerases by generating a
uniform substrate without secondary structure (142). SSBs were also shown to play an important
role in coordinating DNA synthesis on the lagging strand in Bacteria and Eukarya (182, 183), but
this activity has not yet been reported in archaea.
In all domains of life, SSB binds to ssDNA via an oligonucleotide/oligosaccharide binding (OB)
fold (32). In bacteria, the SSB protein forms homotetramers in which each monomer contains a
single OB fold at the N terminus and a flexible acidic C terminus, which is needed for interaction
with other proteins (82).
In Eukarya, the SSB is the heterotrimeric replication protein A (RPA) and comprises the RPA70,
RPA32, and RPA14 proteins (142). RPA70 contains four OB folds, whereas RPA32 and RPA14
each contain one OB fold; however, only three OB folds from RPA70 and the OB fold of RPA32
are involved in ssDNA binding. The RPA70 protein also contains a C4-type zinc-finger motif and
RPA32 contains a WHD, and both are found near the C-terminus. These motifs, however, do
not participate in DNA binding but rather in complex stability and protein-protein interactions
(142).
The single RPA homolog from Methanocaldococcus jannaschii was the first archaeal SSB to be
characterized (70). Other archaeal SSB proteins show significant variation in structure, domain
organization, and subunit composition. In most Aigarchaea and Crenarchaea, the SSBs are similar
to those in Bacteria in that they have a single OB fold at the N terminus followed by a flexible, acidic
C terminus, which is not required for DNA binding (171), whereas the genomes of Korarchaea and
Thaumarchaea contain homologs of both SSB and RPA (143). The structure of the Crenarchaea
OB fold, however, is more similar to that of the eukaryotic RPA70 than to that of the bacterial
SSB (84).
The SSB proteins from the other archaeal lineages are diverse in structure and domain organi-
zation, although overall they are similar to the eukaryotic RPA. Whereas some organisms encode a
single RPA with multiple OB folds (70), others contain multiple RPA homologs (88). It was found
that some organisms with multiple RPAs form trimeric complexes similar to the eukaryotic RPA
complex (88). Genetic studies also showed that in some cases when multiple RPAs are present,
only a subset is essential for cell viability (163). In addition to the OB fold, several archaeal RPA
proteins also contain a zinc-finger motif (147, 163).
In silico approaches identified SSBs in most archaeal species either by similarity to the eu-
karyotic or bacterial SSB or by searches for putative OB folds. However, to date, no canonical
80 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
SSB has been detected in one crenarchaeal order, Thermoproteales. Screening of a Thermoproteus
tenax extract for proteins capable of binding to ssDNA revealed a new type of SSB [referred to as
ThermoDBP (Thermoproteales-specific DNA binding protein)] (136). The protein was shown
to form a homodimer with a novel DNA binding fold at the N terminus and a leucine zipper at
the C terminus that mediates dimerization (136).
The Cdc45-MCM-GINS Complex
In Eukarya, the MCM helicase is not active on its own but is activated by association with two
accessory factors: the GINS complex and the Cdc45 protein. The resulting complex is referred to
as the CMG (Cdc45-MCM-GINS) complex and is thought to function as the replicative helicase
(129). However, in contrast to the eukaryotic Mcm2–7 complex, in vitro studies showed that the
archaeal MCM is active on its own without association with other proteins. Nevertheless, a CMG
complex may also be present in Archaea and may play a role in vivo.
The GINS complex. In Eukarya, the GINS complex is a ring-shaped heterotetramer of four
related proteins, Sld5, Psf1, Psf2, and Psf3 (GINS is derived from the Japanese go-ichi-ni-san,
meaning 5-1-2-3), that plays an essential role in establishment and maintenance of replication forks
(65, 108). In addition to interacting with MCM and Cdc45, the GINS complex also interacts with
the DNA polymerase α(Polα)-primase complex that synthesizes primers on the lagging strand
and with the leading-strand DNA polymerase Polε(65, 108). It is thought that the GINS complex
may be the eukaryotic functional homolog of the bacterial τ-subunit that couples the helicase,
primase, and polymerases at the replication fork and thus plays an important scaffolding role in
replisome formation and in coordinating leading- and lagging-strand synthesis (64).
The first archaeal GINS homologs were identified using in silico approaches that detected
their weak similarity to the eukaryotic proteins (112). This study was followed by the isolation and
characterization of the GINS complex from S. solfataricus and other species (9, 114). Although
some species contain a single homolog, designated GINS15 (also called GINS51) for its similarity
to the eukaryotic Psf1 and Sld5 proteins, other species contain two homologs, a GINS15 and
another referred to as GINS23 for its similarities to the eukaryotic Psf2 and Psf3 proteins (108,
110). However, it has been shown that in all cases studied, the GINS proteins form tetrameric
complexes, either homotetramers of GINS15 (128) or a heterotetramer of two subunits of GINS15
and two of GINS23 (114). As is the case in Eukarya, the archaeal GINS proteins are essential for
cell viability (155).
Supporting evidence for GINS’s function as a scaffolding complex comes from the observation
that the archaeal complex interacts with a number of replisome components, including primase,
MCM, Cdc45, DNA polymerase D (PolD), and PCNA (102, 109, 114, 140). The genes encoding
GINS15 and GINS23 are often in an operon with the genes encoding PCNA and/or the small
subunit of primase (PriS) and MCM, respectively (14, 108).
Structural studies revealed that each of the two GINS subunits is composed of two distinct
domains: a large A domain and a smaller B domain. The order of the two domains, however, is
different in the different subunits. In the GINS15 protein the A domain is at the N terminus and
the B domain is at the C terminus (AB type), whereas the orientation is reversed in GINS23 (BA
type) (65).
The structure of the Thermococcus kodakarensis GINS revealed that two GINS15 proteins form
a top layer and two GINS23 proteins form a bottom layer with an overall structure that resembles
a trapezoid with a narrow cavity in the center (130). The structure suggests that the B domain
of GINS15 does not participate in tetramer formation (130). It is possible that the B domain
www.annualreviews.org Archaeal DNA Replication 81
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
is needed for GINS interaction with other components of the CMG complex or with other
replication proteins.
The structure also revealed similarities between the B domain and the C-terminal part of PriS
(65, 167). This C-terminal domain, however, is not present in PriS from all species. Because the
gene encoding GINS15 in several archaeal species is in an operon with the gene encoding PriS
(14, 108), it has been suggested that the C-terminal domain of PriS may have been acquired from
the GINS15 protein via tandem duplication (167).
Cdc45. The first suggestion that Archaea might contain a Cdc45 homolog came from a study that
identified a nuclease in T. kodakarensis associated with the GINS complex (101, 102). A structural
prediction analysis suggested that the protein, referred to as GAN (GINS-associated nuclease),
has limited similarity to the bacterial RecJ and the eukaryotic Cdc45 proteins (90, 101). Detailed in
silico analyses have shown that all archaeal species contain a protein with similarity to Cdc45 and
the bacterial RecJ (111). Although the T. kodakarensis protein possesses nuclease activity, this is not
the case for all archaeal Cdc45 proteins. When the GINS complex was purified from S. solfataricus,
a protein with similarity to the RecJ DNA binding domain copurified with the complex (114).
This protein does not include the RecJ nuclease domain and does not possess nuclease activity.
Similarly, the eukaryotic Cdc45 protein does not contain key residues for nuclease activity and no
nuclease activity could be detected in vitro with the eukaryotic enzyme ( J. Hurwitz & Z. Kelman,
unpublished results). Therefore, the function of the archaeal Cdc45 in DNA replication probably
does not depend on its nuclease domain. Nevertheless, it is likely that the archaeal and eukaryal
Cdc45 evolved from the bacterial RecJ (110), and the archaeal Cdc45 and the bacterial RecJ have
similar three-dimensional structures (S. Nair & Z. Kelman, unpublished results).
The role of Cdc45 in archaeal DNA replication is not clear. The protein interacts with GINS
and PolD, which may suggest a role at the replication fork. However, although the eukaryotic
Cdc45 is essential for cell viability (115), the genes encoding the archaeal enzyme can be deleted
from the chromosome without a major effect on cell growth (T. Santangelo & Z. Kelman, un-
published results). This may suggest that the protein is not essential for DNA replication or that
another factor can replace it. Future studies are needed to determine the role, if any, for the
archaeal Cdc45 protein in vivo.
ELONGATION COMPLEX
Following MCM activation, the replication bubble is formed, and the other components of the
replisome are recruited to the DNA to establish the two replication forks and initiate bidirectional
DNA synthesis.
Primase
DNA polymerases are incapable of initiating DNA synthesis de novo and require a 3-hydroxyl-
primed template in order to elongate DNA chains. DNA primases synthesize short RNA primers
on template DNA, which are subsequently extended by DNA polymerase (Figure 2;Figure 3a)
(44). In bacteria, DNA primase consists of a single subunit, the DnaG protein, which associates
with both the DnaB helicase and PolIII. It synthesizes a 10–12 oligoribonucleotide primer
that is transferred to the bacterial clamp loader. In Eukarya, DNA primase is a heterodimer
containing a catalytic p48 subunit in tight association with a regulatory p58 subunit. In vivo, the
heterodimer is found in a complex with two other proteins, the Polαcatalytic subunit (p180)
and the B subunit (p70), to form the Polα-primase complex (44). The primase component
82 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
synthesizes 10–15 oligoribonucleotide primers that are elongated by the Polαcomponent,
generating covalently linked RNA-DNA oligonucleotide chains of 30–40 nucleotides (44). These
pre-Okazaki fragments are recognized by the clamp loader.
The archaeal primase is a two-subunit complex; a small subunit p41 (PriS) contains the catalytic
activity, and a large subunit p46 (PriL) regulates primase activity and contains an iron-sulfur
domain. Both subunits were shown to be essential for cell viability (155). Nanoarchaeota genomes
encode a shorter form of primase in which the iron-sulfur domain of PriL is fused to the catalytic
domain of PriS (143).
The first report on the biochemical properties of the archaeal primase came from studies
on the small subunit from M. jannaschii (31), followed by studies on the dimeric enzyme from
other species (60). No homologs of the eukaryotic Polαor B subunit have been identified in
archaeal genomes (96). In contrast to DNA primases from Bacteria and Eukarya that can utilize
only ribonucleotides, in vitro studies with the archaeal catalytic subunit PriS and the two-subunit
complex PriS-PriL demonstrated that the archaeal primase is capable of initiating oligonucleotide
chains de novo from either ribonucleotide triphosphates (rNTPs) or deoxyribonucleotide
triphosphates (dNTPs) (60, 96).
Although the primase can utilize both ribo- and deoxynucleotides in vitro, several different
assays, including RNA unmasking, RNA labeling, and RIP mapping, strongly suggest that primers
are composed of oligoribonucleotides in vivo (119, 149). The much higher cellular levels of rNTPs
in comparison to dNTPs (89) may explain the discrepancies between primase activity in vitro and
primer synthesis in vivo.
Archaeal primases also have strand displacement, gap-filling, terminal transferase, and py-
rophosphatase activity (60), but the roles of those activities are not clear. In vitro studies also
showed that the enzymes are capable of generating nucleotide adducts in the presence of rNTPs,
dNTPs, and small molecules with OH and NH groups. The reaction results in the formation
of r/dNMP-O-R and r/dNMP-N-R (19, 20). It is not clear whether the enzyme makes similar
compounds in vivo and whether such products have a physiological role.
The three-dimensional structure of PriS from several species has been determined, revealing
a two-domain structure (95, 137): a large primase domain (62), which contains the catalytic part,
and a small domain that appears to have a species-specific fold. The catalytic domain also contains
a zinc-finger motif. Although the role of zinc binding is not clear, the proximity of this motif
to the active site may suggest a role in catalysis. DnaG also contains a zinc binding domain that
participates in ssDNA binding (25), and therefore it is possible that zinc binding has a similar role
in the archaeal primase.
The structures of the N-terminal part of PriL alone and a complex of the N-terminal part of
PriL and PriS have been determined (95). The N-terminal part of PriL interacts with PriS and
also forms an arm that connects the catalytic subunit to the DNA binding domain located at the
C terminus of PriL (116). The C-terminal part of PriL contains an iron-sulfur domain (87). The
structure and role of the iron-sulfur domain is not yet known, but this domain may play a role in
maintaining the correct three-dimensional structure of the C-terminal domain (137).
Polymerase Accessory Proteins
The replicative polymerase on its own has very low processivity. High processivity is conferred
by a ring-shaped processivity factor (sliding clamp) that encircles DNA and acts to tether the
polymerase catalytic unit to the template for processive DNA synthesis (Figure 2;Figure 3a) (58,
94, 131). The sliding clamps are stable rings and thus cannot assemble themselves around DNA
but must be loaded onto DNA by a clamp loader complex (69, 181). The clamp loader recognizes
www.annualreviews.org Archaeal DNA Replication 83
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
the 3end of the single strand–duplex (primer-template) junction and utilizes ATP hydrolysis
to assemble the clamp around the primer. The clamp encircles the primer and then binds the
polymerase for rapid and processive DNA synthesis. Upon completion of an Okazaki fragment,
the polymerase dissociates from the clamp, leaving it assembled around the duplex DNA. The
clamps left on the DNA interact with proteins needed for Okazaki fragment maturation, chromatin
remodeling, and other cellular processes (170).
Replication factor C. In eukarya, the clamp loader is the pentameric replication factor C (RFC)
complex composed of one large subunit (Rfc1) and four small subunits (Rfc2-5), all belonging
to the AAA+family of ATPases (69, 181). Most archaea contain two homologs of RFC (22).
One is similar to the small subunits of the eukaryotic RFC and thus is referred to as RFC small
(RFCS), and the other is similar to the large subunit of the eukaryotic RFC and is named RFC
large (RFCL). Like the eukaryotic RFC complex, the archaeal complex is pentameric, with four
subunits of RFCS associated with one subunit of RFCL (58, 127).
RFC must be bound to ATP in order to associate with and open the PCNA ring. Following
the assembly of PCNA around the primer, the interaction of RFC with the DNA and PCNA
stimulates its ATPase activity. Upon ATP hydrolysis, the affinity of RFC to DNA and PCNA is
substantially reduced, resulting in the dissociation of RFC from the DNA and PCNA (58, 127).
In Bacteria and Eukarya, primase remains associated with the primer following primer synthesis.
It was shown that primases and the clamp loaders interact with SSB, and competition between the
clamp loader and primase for the interaction with SSB plays an important role in the handoff of the
primer from primase to the clamp loader (182, 183). A similar RPA-dependent handoff mechanism
may also exist in archaea, although it has not yet been reported. Studies have shown, however,
that the archaeal primase interacts directly with RFC (109, 178). It is thus possible that in Archaea
the handoff of the primers from primase to RFC does not require the involvement of RPA.
As mentioned above, PCNA remains on the DNA after the completion of an Okazaki fragment.
However, during replication there are more Okazaki fragments than PCNA trimers within the
cell, and so the clamps must be recycled. It was shown that the eukaryotic RFC can actively remove
PCNA rings from DNA (180) and a similar unloading activity was proposed for archaeal RFC (77).
Biochemical, biophysical, structural, and molecular modeling studies with RFC have shed light
on the mechanism by which RFC assembles the PCNA around DNA (68, 124). The data suggest
that during the assembly process, RFC, in the ATP-bound form, forms a right-handed spiral on
the top of the clamp with a pitch that is congruent with the helical geometry of duplex DNA.
Binding to the clamp results in clamp opening with an out-of-plane configuration resembling a
right-handed spring washer (69), enabling assembly around the DNA.
Proliferating cell nuclear antigen. PCNA is a trimeric ring that encircles dsDNA and can slide
bidirectionally along it. All activities described for the PCNA proteins require them to encircle
the DNA duplex; no biochemical function for PCNA separate from DNA has been reported.
PCNA was first reported as a processivity factor for replicative DNA polymerases, but subsequent
studies have established that PCNAs also associate with, and modulate the activity of, many other
proteins involved in nucleic acid metabolic transactions and cell-cycle regulation (131, 170). It
was proposed that the movement of PCNA along dsDNA might function as a moving platform
for enzymes that participate in DNA metabolic processes but have low affinity for DNA (76).
Most of the proteins that interact with PCNA do so via a PIP motif (173). The PIP motif
interacts with the loop that connects the two domains in each PCNA monomer [referred to as
the interdomain connecting loop (IDCL)] (49). Biochemical and structural analysis elucidated the
requirement for an intact PIP motif for interaction with PCNA (60).
84 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
The three-dimensional structures of PCNA proteins from several archaeal species were de-
termined (131), as were PCNA complexes with interacting enzymes, including DNA ligase, flap
endonuclease 1 (Fen1), and DNA polymerase B (PolB) (60). Although the crenarchaeota PCNA
protein forms heterotrimers while PCNA from the other kingdoms form homotrimers, the ring
structures are very similar (131), and they are also similar to the bacterial β-subunit and the
eukaryotic PCNA (58, 80).
The structure revealed that three monomers interact in a head-to-tail manner to form a trimeric
ring sufficiently large to accommodate dsDNA. Each monomer is composed of two structurally
similar domains, so the trimer has a pseudo six-fold symmetry (75). The two domains in each
monomer are connected by the IDCL (49).
Although all PCNA proteins are acidic (80), the charge distribution on the ring surface is not
symmetrical. The outer surface is negatively charged, whereas there is a net positive electrostatic
potential in the central cavity where the dsDNA is located (131). The positive charge in the
central cavity was shown to be required for interaction with the DNA, and it was suggested that
the negatively charged surface might prevent nonspecific interaction with DNA. In contrast to the
structures of other PCNA proteins, the structures of the halophilic PCNA proteins revealed very
few positively charged residues within the central cavity (125, 176), which may suggest a different
mechanism of interaction for halophile PCNA with DNA (125).
When the structures of the sliding clamps were initially determined, it was hypothesized that the
DNA passes perpendicularly through the central hole, allowing bidirectional sliding along dsDNA
(180). However, structural studies of PCNA-DNA complexes from several archaeal species suggest
that the DNA has a substantial tilt while passing through the central hole of the ring (121).
Several mechanisms regulate the activity of eukaryotic PCNA. One involves the binding of
small PIP-containing regulatory proteins to PCNA, which in turn prevents other proteins from
binding and thus regulates (inhibits) the effect of PCNA on their activity (41). Archaea may
use similar mechanisms in which binding of small proteins to PCNA regulates its interaction
with other enzymes. Studies have shown that a small protein from T. kodakarensis,referredtoas
Thermococcales inhibitor of PCNA (TIP), binds to PCNA and prevents its interaction with DNA
polymerase and other enzymes. However, the mechanism of binding of TIP to PCNA does not
involve a canonical PIP motif (99).
Another mechanism by which the eukaryotic PCNA is regulated involves modification by
small proteins, such as ubiquitin and SUMO. These proteins, when binding PCNA, modulate
PCNA interaction with other enzymes (33). Although small modifier proteins have also been
identified in Archaea (120), it remains to be determined whether they modify PCNA and whether
the modification plays a role in regulating PCNA function.
A third mechanism used by Eukarya to regulate PCNA activity is phosphorylation of the
interacting enzymes, which results in the modulation of their interactions with PCNA (56). Protein
phosphorylation is common in Archaea (83), and in vivo phosphorylation of PCNA and several
replication proteins has been reported (144). Therefore, it is possible that phosphorylation may
regulate the activity of the archaeal PCNA.
DNA Polymerase
DNA replication is achieved by a DNA-dependent DNA polymerase that uses primed ssDNA as a
template to synthesize the complementary strand. Bacterial PolIII replicates both the leading and
lagging strands (127). In Eukarya, two different polymerases, Polεand Polδ, replicate the leading
and lagging strands, respectively (127). Two different DNA polymerases, PolB and PolD, have
been implicated in archaeal DNA replication.
www.annualreviews.org Archaeal DNA Replication 85
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
B-type polymerases. The archaeal PolB DNA polymerases have been extensively studied, mainly
for their importance in polymerase chain reaction. A large number of enzymes have been identified,
purified, and biochemically characterized (60, 110, 139). At least one PolB homolog has been
identified in every archaeal genome. In addition to polymerase activity, these enzymes possess a
3to 5exonuclease proofreading activity. Polymerases are not processive; processivity is achieved
by PolB interaction with PCNA.
For extended processivity, bacterial and eukaryal replicative polymerases require SSB. In vitro
studies showed, however, that at least in some archaea, RPA inhibited rather than stimulated the
primer extension activity of PolB (81), and direct interaction between RPA and PolB has also
been reported (81, 102). The presence of PCNA relieved, but did not eliminate, the inhibitory
effect of RPA. Why RPA inhibits PolB and what the role of the inhibition is in vivo is not clear.
It is possible that when a polymerase is a part of the replisome, RPA does not inhibit its activity.
Alternatively, PolB may not be the replicative enzyme in Archaea (29, 155) but instead is involved
in DNA repair and/or recombination. If this is the case, then inhibition by RPA may insure that
PolB will replicate only short DNA fragments, as expected for a repair enzyme. It is also possible
that PolB replicates only the leading strand, and therefore its ssDNA template is not coated with
RPA. Future studies are needed to determine how common the inhibition of PolB by RPA is, as
well as the physiological role of that inhibition.
The three-dimensional structures of several archaeal PolB enzymes have been determined
and revealed similar topology to other DNA polymerases, including the palm, finger, thumb,
and exonuclease domains (60, 73). One feature of the archaeal PolB not found in other family
members is an N-terminal domain involved in specific interactions with uracil and hypoxanthine
on the template strand (24). Copying of damaged DNA could result in irreversible mutations.
Therefore, the ability of a polymerase to recognize uracil or hypoxanthine, followed by replication
stalling, may enable the cell to repair the DNA damage prior to replication (24, 60).
D-type polymerases. The genomes of all Archaea except Crenarchaea contain the archaeal-
specific PolD in addition to PolB. PolD was originally isolated from Pyrococcus furiosus by screening
for DNA polymerase activity in cell extract (169). PolD is a dimeric enzyme, and the large subunit
(DP2) contains polymerase catalytic activity while the small subunit (DP1) contains the 3to 5
exonuclease activity. The DP2 protein does not display any sequence similarity with other protein
families (110), whereas DP1 shares amino acid sequence similarity with several of the small,
noncatalytic subunits of the eukaryotic Polα,Polδ,andPolε(1). Although each subunit alone has
low activity, the dimeric PolD complex exhibits polymerase and exonuclease activities (60).
The three-dimensional structures of the N-terminal parts of DP1 and DP2 have been solved
(117, 179). The structure of the N-terminal part of DP1 confirmed the bioinformatic prediction
that it shares similarity with the small subunits of Polαand Polε(179). Although high-resolution
structures of the full-length proteins have not yet been reported, bioinformatic predictions provide
insight into the structure and domain organization, including the presence of an OB fold in DP1
and a zinc-finger motif in DP2 (110). Mutational analyses of conserved residues, regions, and
motifs have shed light on the regions required for activity, on interactions between DP1 and DP2,
and on interactions with DNA (157).
Which is the replicative polymerase in Archaea? The three replicative polymerases in Eukarya,
Polα,Polδ,andPolε, all belong to family B, and therefore it was presumed that members of
this family also replicate the archaeal genomes. Crenarchaea genomes encode only members of
PolB, and therefore PolB must be the replicative enzyme in this kingdom. In other archaeal
branches, however, both PolB and PolD are present. It is not clear whether both polymerases
are involved in chromosomal replication, as in Eukarya, or only one of the two is, as in Bacteria,
86 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
with the other polymerase playing a different cellular role. Genetic studies with PolB from several
species suggested that both PolB and PolD are required for cell viability (13). In other organisms,
however, PolB is dispensable for cell growth (29, 155). These contradictory observations need
further evaluation. They may suggest, however, that at least in some species, PolD replicates both
the leading and lagging strands.
Aphidicolin is an inhibitor of the eukaryotic replicative polymerases and thus was thought to
inhibit all members of the family B enzymes. PolD, however, is not sensitive to aphidicolin (48,
169). Therefore, sensitivity to aphidicolin was suggested to serve as an additional tool to determine
whether the archaeal PolB is essential for cell growth. The picture, however, is not that clear. In
vitro studies have shown that although aphidicolin inhibits PolB from some species (48, 169), it
has no effect on PolB activity in others (81). In vivo studies did not provide much clarification.
Although aphidicolin was shown to inhibit the growth of Halobacterium halobium, there was no
effect on the growth of the crenarchaea Sulfolobus acidocaldarius (43). The S. acidocaldarius genome
contains only PolB, again demonstrating that some archaeal PolB enzymes are not sensitive to
the drug. Taking the in vivo and in vitro evidence together, it is clear that aphidicolin sensitivity
cannot be used as a tool to determine whether PolB is required for chromosomal replication.
Additional studies are needed to determine whether both PolB and PolD are a part of the
replisome or whether only PolD is involved in chromosomal replication and PolB has another
function. It is possible that although both polymerases are replicative enzymes, in the absence
of PolB, PolD can substitute for PolB activity. Although in Eukarya Polεis the leading-strand
polymerase (127), it was shown that cells harboring mutant Polεare viable and that Polδcan
replicate both DNA strands (85). It is possible that in archaea, PolD replicates the lagging strand
and PolB copies the leading strand. The role for PolD in lagging-strand synthesis may be supported
by the observation that PolD exhibits a strand displacement activity (54) that is required for Okazaki
fragment maturation. In addition, PolD can utilize RNA primers more efficiently than PolB (55),
whereas processivity of PolB, in the presence of PCNA, is greater than that of PolD (132). Similar
observations were made in Eukarya, where the leading-strand polymerase is more processive than
the lagging-strand enzyme (12). In addition, the inhibition of PolB by RPA may suggest that it
is not replicating the lagging strand because the lagging-strand polymerase probably needs to
displace RPA.
OKAZAKI FRAGMENT MATURATION
During chromosomal replication, the leading strand is replicated as a continuous strand, whereas
the lagging strand is replicated discontinuously as a series of Okazaki fragments (Figure 3). The
RNA primers that initiate each Okazaki fragment must be removed, the gap in the DNA must be
filled, and the newly synthesized Okazaki fragment must be ligated to the previous fragment to
create the mature dsDNA. This task is achieved by the concerted activity of several enzymes, in-
cluding DNA polymerase, Fen1, and DNA ligase (Figure 3b) (5). The lagging-strand polymerase,
upon reaching the previous Okazaki fragment, displaces the primer and provides the substrate for
Fen1. Following Fen1 removal of the primer, the lagging-strand DNA polymerase fills the gap
and DNA ligase joins the two adjacent Okazaki fragments.
Flap Endonuclease 1
Fen1 is a structure-specific nuclease that plays an important role in Okazaki fragment maturation
on the lagging strand (4). The lagging-strand DNA polymerase, using its strand displacement
activity, removes the RNA primer and forms a branched DNA molecule (flap structure). The
5-flap endonuclease activity of Fen1 removes the primer (Figure 3b).
www.annualreviews.org Archaeal DNA Replication 87
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
The three-dimensional structures of Fen1 from several archaeal species and co-complexes
with DNA have been determined (60). The structures reveal that the protein catalytic domain is
similar to other nucleases. In addition, Fen1 contains a helical clamp that encircles the 5-flap and
interacts with the upstream and downstream duplex DNA. Upon binding to the substrate, the
protein undergoes a conformational change that closes the clamp around the single-stranded flap.
In Eukarya, it was shown that the preferred substrate of Fen1 contains an unpaired 3nucleotide
(3-flap) overlapping with a variable length region of 5ssDNA (5-flap), referred to as an overlap-
flap (39). The structure of the archaeal Fen1 with DNA suggested that an overlap-flap may also
be the preferred substrate for the archaeal enzyme (18).
As is the case with the eukaryotic Fen1, the archaeal enzyme is not essential for cell viability
(155). In Eukarya, Exonuclease I (ExoI), a homolog of Fen1 with similar catalytic activity, may
replace Fen1 activity in the deletion strain (4). However, in Archaea there are no clear homologs
of ExoI or other known nucleases with Fen1-like activity. Thus, it is not clear which protein(s)
replaces Fen1 activity in the deleted strain, although it was suggested that the archaeal Cdc45 might
substitute for Fen1 in the deleted strain (101). Support for this idea comes from the observation that
although Cdc45 or Fen1 can be readily deleted from T. kodakarensis cells, the double mutation is
lethal (T. Santangelo & Z. Kelman, unpublished results). Alternatively, RNase H, which degrades
RNA in a RNA-DNA hybrid, may be responsible for primer removal in the Fen1 deleted strain.
Future studies are needed, however, to identify the mechanism by which Okazaki fragments mature
in the absence of Fen1.
DNA Ligase
DNA ligase plays an essential role in chromosomal DNA replication by joining adjacent Okazaki
fragments on the lagging strand. DNA ligases can be divided into two groups according to the
cofactor required for activity: ATP or NAD+.NAD
+-dependent ligases are found predominantly
in Bacteria, whereas eukaryotic genomes contain ATP-dependent ligases (110).
Genes that encode DNA ligase have been identified in all archaeal genomes and were shown
to be essential for cell viability. The first archaeal DNA ligase was identified in the genome of the
thermophilic archaeon Desulfolobus ambivalens (86). The first characterized archaeal ligase from
M. thermautotrophicus demonstrated a strict requirement for ATP as a cofactor (165), but subse-
quent studies identified a small subset of archaeal ligases that are NAD+-dependent (185), whereas
others can utilize both cofactors (60).
The structures of several archaeal DNA ligases were determined (121, 135) and together
with the structure of the eukaryotic DNA ligase bound to DNA (134) revealed a three-domain
organization consisting of a DNA binding domain (DBD), a catalytic nucleotidyltransferase do-
main, and a C-terminal OB fold domain. In the absence of DNA substrate, the enzyme has an
open, extended form. Upon DNA binding, the enzyme forms a closed structure around the ds-
DNA with both the DBD and OB domains binding to the minor groove on both sides of the
nick, thus positioning the catalytic domain for catalysis. It was suggested that binding to PCNA
facilitates the transition from extended to closed structure (57).
The Role of Proliferating Cell Nuclear Antigen in
Okazaki Fragment Maturation
The enzymes required for Okazaki fragment maturation, including DNA polymerase, Fen1, and
DNA ligase, interact with PCNA via a PIP motif, and interaction with PCNA modulates their
enzymatic activities (60, 170). These interactions suggest that PCNA coordinates the Okazaki
fragment maturation process by sequentially recruiting factors to the lagging strand.
88 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
In crenarchaea, PCNA is a heterotrimer and each subunit of PCNA has specificity for binding to
DNA polymerase, Fen1, or DNA ligase (8). This specificity is likely to have arisen from differences
in the structure of the IDCL among the three different PCNA proteins (175), enabling them to
discriminate between small differences in the PIP motifs of the different interacting enzymes. The
simultaneous binding of all three proteins to PCNA provides an efficient coupling of the three
enzyme activities. Upon flap formation by a lagging-strand polymerase, Fen1 cleaves the flap, the
gap is filled by the polymerase, and the Okazaki fragments are joined by DNA ligase.
PCNA proteins from all other archaeal species, however, form homotrimeric rings (131).
Therefore, it is not clear whether the role of PCNA in regulating Okazaki fragment maturation
in these species is similar to that in Crenarchaea. Although DNA polymerase, Fen1, and DNA
ligase all interact with PCNA, it is not clear whether they can bind simultaneously.
TERMINATION OF DNA REPLICATION AND
PRODUCT RESOLUTION
At the end of chromosomal replication, converging replication forks must complete replication
and the replisomes must dissociate from the DNA. It is essential that converging forks not pass each
other and continue replication, as this results in over-replication of part of the chromosome. In
Bacteria, termination occurs in a specific chromosomal region referred to as the termination region
(ter). The specific ter-binding protein Tus participates in the termination process by regulating
replisome movement, ensuring precise replication of the chromosome (146).
Limited information is available on the mechanism of replication termination in Archaea. The
circular nature of archaeal chromosomes may suggest a mechanism similar to bacterial termination,
but a study using S. solfataricus suggested that termination occurs by random collision of the two
replication forks and not at a specific site (37). In addition, the presence of ter-like sequences or
Tus-like proteins has not yet been reported in Archaea.
Bacterial chromosomes contain dif (deletion induced filamentation) regions near ter,where
concatemers formed during chromosome segregation are resolved by the Xer site-specific re-
combinases (XerC and XerD) in conjunction with FtsK (146). Only a few studies on archaeal
chromosome resolution have been published. Archaeal genomes contain a dif-like sequence and
Xer homologs (26, 37), but the location of the dif sequence in the chromosome is different in
different species. In some species the dif region is located in the termination zone (26), but it is
not near the termination region in others (37). The results to date, although limited, suggest that
Archaea may use a bacterial-like mechanism for chromosome resolution.
CONCLUDING REMARKS
Much progress has been made in understanding archaeal DNA replication, but there remains
much to be elucidated. Many factors have been identified and characterized, and the three-
dimensional structures of most proteins have been determined. New genetic tools (38) have
identified new proteins that may be involved in replication. These genetic tools, together with
fluorescence and single-molecule approaches, should lead to new and exciting studies on the
replication process in vivo.
Future studies will concentrate on poorly understood aspects of the replication process. The
mechanism that regulates the initiation process is unknown, and the coordination between ini-
tiation and other cell-cycle events is yet to be explored. Similarly, the study of termination and
chromosome segregation is in its infancy and more research is needed.
The development of an in vitro replication system is one of the goals of future research.
Replication in bacteria, viruses, and phages was characterized after in vitro replication systems
www.annualreviews.org Archaeal DNA Replication 89
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
were developed, and it is anticipated that such a system will similarly serve as a catalyst for the
understanding of archaeal replication.
DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
We thank Drs. Jerard Hurwitz, Satish Nair, and Thomas Santangelo for sharing data prior to
publication. We thank Dr. Debra Weinstein for help with figures. We wish to apologize to
colleagues whose primary work was not cited because of space limitations.
LITERATURE CITED
1. Aravind L, Koonin EV. 1998. Phosphoesterase domains associated with DNA polymerases of diverse
origins. Nucleic Acids Res. 26:3746–52
2. Aravind L, Koonin EV. 1999. DNA-binding proteins and evolution of transcription regulation in the
archaea. Nucleic Acids Res. 27:4658–70
3. Arora J, Goswami K, Saha S. 2014. Characterization of the replication initiator Orc1/Cdc6 from the
archaeon Picrophilus torridus.J. Bacteriol. 196:276–86
4. Balakrishnan L, Bambara RA. 2013. Flap endonuclease 1. Annu. Rev. Biochem. 82:119–38
5. Balakrishnan L, Bambara RA. 2013. Okazaki fragment metabolism. Cold Spring Harb. Perspect. Biol.
5:a010173
6. Barry ER, Lovett JE, Costa A, Lea SM, Bell SD. 2009. Intersubunit allosteric communication mediated
by a conserved loop in the MCM helicase. Proc. Natl. Acad. Sci. USA 106:1051–56
7. Barry ER, McGeoch AT, Kelman Z, Bell SD. 2007. Archaeal MCM has separable processivity, substrate
choice and helicase domains. Nucleic Acids Res. 35:988–98
8. Beattie TR, Bell SD. 2012. Coordination of multiple enzyme activities by a single PCNA in archaeal
Okazaki fragment maturation. EMBO J. 31:1556–67
9. Bell SD. 2011. DNA replication: archaeal oriGINS. BMC Biol. 9:36
10. Bell SD. 2012. Archaeal orc1/cdc6 proteins. Subcell. Biochem. 62:59–69
11. Bell SD, Botchan MR. 2013. The minichromosome maintenance replicative helicase. Cold Spring Harb.
Perspect. Biol. 5:a012807
12. Bermudez VP, Farina A, Raghavan V, Tappin I, Hurwitz J. 2011. Studies on human DNA polymerase
εand GINS complex and their role in DNA replication. J. Biol. Chem. 286:28963–77
13. Berquist BR, DasSarma P, DasSarma S. 2007. Essential and non-essential DNA replication genes in the
model halophilic Archaeon, Halobacterium sp. NRC-1. BMC Genet. 8:31
14. Berthon J, Cortez D, Forterre P. 2008. Genomic context analysis in Archaea suggests previously unrec-
ognized links between DNA replication and translation. Genome Biol. 9:R71
15. Boulikas T. 1996. Common structural features of replication origins in all life forms. J. Cell Biochem.
60:297–316
16. Brewster AS, Chen XS. 2010. Insights into the MCM functional mechanism: lessons learned from the
archaeal MCM complex. Crit. Rev. Biochem. Mol. Biol. 45:243–56
17. Brewster AS, Slaymaker IM, Afif SA, Chen XS. 2010. Mutational analysis of an archaeal minichromosome
maintenance protein exterior hairpin reveals critical residues for helicase activity and DNA binding. BMC
Mol. Biol. 11:62
18. Chapados BR, Hosfield DJ, Han S, Qiu J, Yelent B, et al. 2004. Structural basis for FEN-1 substrate
specificity and PCNA-mediated activation in DNA replication and repair. Cell 116:39–50
90 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
19. Chemnitz Galal W, Pan M, Giulian G, Yuan W, Li S, et al. 2012. Formation of dAMP-glycerol and
dAMP-Tris derivatives by the Thermococcus kodakaraensis DNA primase. J. Biol. Chem. 287:16220–29
20. Chemnitz Galal W, Pan M, Kelman Z, Hurwitz J. 2012. Characterization of the DNA primase complex
isolated from the archaeon, Thermococcus kodakaraensis.J. Biol. Chem. 287:16209–19
21. Chen SH, Chan N-L, Hsieh T-S. 2013. New mechanistic and functional insights into DNA topoiso-
merases. Annu. Rev. Biochem. 82:139–70
22. Chia N, Cann I, Olsen GJ. 2010. Evolution of DNA replication protein complexes in eukaryotes and
archaea. PLoS ONE 5:e10866
23. Chong JP, Hayashi MK, Simon MN, Xu RM, Stillman B. 2000. A double-hexamer archaeal minichro-
mosome maintenance protein is an ATP-dependent DNA helicase. Proc. Natl. Acad. Sci. USA 97:1530–35
24. Connolly BA. 2009. Recognition of deaminated bases by archaeal family-B DNA polymerases. Biochem.
Soc. Trans. 37:65–68
25. Corn JE, Pease PJ, Hura GL, Berger JM. 2005. Crosstalk between primase subunits can act to regulate
primer synthesis in trans. Mol. Cell 20:391–401
26. Cortez D, Quevillon-Cheruel S, Gribaldo S, Desnoues N, Sezonov G, et al. 2010. Evidence for a Xer/dif
system for chromosome resolution in archaea. PLoS Genet. 6:e1001166
27. Costa A, Hood IV, Berger JM. 2013. Mechanisms for initiating cellular DNA replication. Annu. Rev.
Biochem. 82:25–54
28. Costa A, Onesti S. 2009. Structural biology of MCM helicases. Crit. Rev. Biochem. Mol. Biol. 44:326–42
29. ˇ
Cubo ˇ
nov´
a L, Richardson T, Burkhart BW, Kelman Z, Reeve JN, et al. 2013. Archaeal DNA polymerase
D but not DNA polymerase B is required for genome replication in Thermococcus kodakarensis.J. Bacteriol.
195:2322–28
30. De Felice M, Esposito L, Pucci B, De Falco M, Manco G, et al. 2004. Modular organization of a Cdc6-like
protein from the crenarchaeon Sulfolobus solfataricus.Biochem. J. 381:645–53
31. Desogus G, Onesti S, Brick P, Rossi M, Pisani FM. 1999. Identification and characterization of a DNA
primase from the hyperthermophilic archaeon Methanococcus jannaschii.Nucleic Acids Res. 27:4444–50
32. Dickey TH, Altschuler SE, Wuttke DS. 2013. Single-stranded DNA-binding proteins: multiple domains
for multiple functions. Structure 21:1074–84
33. Dieckman LM, Freudenthal BD, Washington MT. 2012. PCNA structure and function: insights from
structures of PCNA complexes and post-translationally modified PCNA. Subcell. Biochem. 62:281–99
34. Duderstadt KE, Berger JM. 2008. AAA+ATPasesin the initiation o f DNA replication. Crit. Rev. Biochem.
Mol. Biol. 43:163–87
35. Duderstadt KE, Berger JM. 2012. A structural framework for replication origin opening by AAA+
initiation factors. Curr. Opin. Struct. Biol. 23:144–53
36. Dueber EL, Corn JE, Bell SD, Berger JM. 2007. Replication origin recognition and deformation by a
heterodimeric archaeal Orc1 complex. Science 317:1210–13
37. Duggin IG, Dubarry N, Bell SD. 2011. Replication termination and chromosome dimer resolution in
the archaeon Sulfolobus solfataricus.EMBO J. 30:145–53
38. Farkas JA, Picking JW, Santangelo TJ. 2013. Genetic techniques for the archaea. Annu. Rev. Genet.
47:539–61
39. Finger LD, Atack JM, Tsutakawa S, Classen S, Tainer J, et al. 2012. The wonders of flap endonucleases:
structure, function, mechanism and regulation. Subcell. Biochem. 62:301–26
40. Fletcher RJ, Bishop BE, Leon RP, Sclafani RA, Ogata CM, Chen XS. 2003. The structure and function
of MCM from archaeal M. thermoautotrophicum.Nat. Struct. Biol. 10:160–67
41. Flores-Rozas H, Kelman Z, Dean FB, Pan ZQ, Harper JW, et al. 1994. Cdk-interacting protein 1
directly binds with proliferating cell nuclear antigen and inhibits DNA replication catalyzed by the
DNA polymerase δholoenzyme. Proc. Natl. Acad. Sci. USA 91:8655–59
42. Forterre P. 2012. Introduction and historical perspective. In DNA Topoisomerases and Cancer,ed.
Y Pommier, pp. 1–52. New York: Humana Press
43. Forterre P, Elie C, Kohiyama M. 1984. Aphidicolin inhibits growth and DNA synthesis in halophilic
archaebacteria. J. Bacteriol. 159:800–2
44. Frick DN, Richardson CC. 2001. DNA primases. Annu. Rev. Biochem. 70:39–80
www.annualreviews.org Archaeal DNA Replication 91
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
45. Gaudier M, Schuwirth BS, Westcott SL, Wigley DB. 2007. Structural basis of DNA replication origin
recognition by an ORC protein. Science 317:1213–16
46. Grabowski B, Kelman Z. 2001. Autophosphorylation of the archaeal Cdc6 homologues is regulated by
DNA. J. Bacteriol. 183:5459–64
47. Grabowski B, Kelman Z. 2003. Archaeal DNA replication: eukaryal proteins in a bacterial context. Annu.
Rev. Microbiol. 57:487–516
48. Greenough L, Menin JF, Desai NS, Kelman Z, Gardner AF. 2014. Characterization of Family D DNA
polymerase from Thermococcus sp. 9N. Extremophiles 18:653–64
49. Gulbis JM, Kelman Z, Hurwitz J, O’Donnell M, Kuriyan J. 1996. Structure of the C-terminal region of
p21WAF1/CIP1 complexed with human PCNA. Cell 87:297–306
50. Guy L, Saw JH, Ettema TJG. 2014. The archaeal legacy of eukaryotes: a phylogenomic perspective.
Cold Spring Harb. Perspect. Biol. In press
51. Haugland GT, Sakakibara N, Pey AL, Rollor CR, Birkeland N-K, Kelman Z. 2008. Thermoplasma
acidophilum Cdc6 protein stimulates MCM helicase activity by regulating its ATPase activity. Nucleic
Acids Res. 36:5602–9
52. Haugland GT, Shin J-H, Birkeland NK, Kelman Z. 2006. Stimulation of MCM helicase activity by a
Cdc6 protein in the archaeon Thermoplasma acidophilum.Nucleic Acids Res. 34:6337–44
53. Hawkins M, Malla S, Blythe MJ, Nieduszynski CA, Allers T. 2013. Accelerated growth in the absence
of DNA replication origins. Nature 503:544–47
54. Henneke G. 2012. In vitro reconstitution of RNA primer removal in Archaea reveals the existence of
two pathways. Biochem. J. 447:271–80
55. Henneke G, Flament D, Hubscher U, Querellou J, Raffin JP. 2005. The hyperthermophilic eur-
yarchaeota Pyrococcus abyssi likely requires the two DNA polymerases D and B for DNA replication.
J. Mol. Biol. 350:53–64
56. Henneke G, Koundrioukoff S, Hubscher U. 2003. Phosphorylation of human Fen1 by cyclin-dependent
kinase modulates its role in replication fork regulation. Oncogene 22:4301–13
57. Howes TR, Tomkinson AE. 2012. DNA ligase I, the replicative DNA ligase. Subcell. Biochem. 62:327–41
58. Indiani C, O’Donnell M. 2006. The replication clamp-loading machine at work in the three domains of
life. Nat. Rev. Mol. Cell Biol. 7:751–61
59. Ishino S, Fujino S, Tomita H, Ogino H, Takao K, et al. 2011. Biochemical and genetical analyses
of the three MCM genes from the hyperthermophilic archaeon, Thermococcus kodakarensis.Genes Cells
16:1176–89
60. Ishino Y, Ishino S. 2012. Rapid progress of DNA replication studies in Archaea, the third domain of life.
Sci. China Life Sci. 55:386–403
61. Ishino S, Kelman LM, Kelman Z, Ishino Y. 2013. The archaeal DNA replication machinery: past, present
and future. Genes Genet. Syst. 88:315–19
62. Iyer LM, Koonin EV, Leipe DD, Aravind L. 2005. Origin and evolution of the archaeo-eukaryotic
primase superfamily and related palm-domain proteins: structural insights and new members. Nucleic
Acids Res. 33:3875–96
63. Jenkinson ER, Chong JP. 2006. Minichromosome maintenance helicase activity is controlled by N- and
C-terminal motifs and requires the ATPase domain helix-2 insert. Proc. Natl. Acad. Sci. USA 103:7613–18
64. Johnson A, O’Donnell M. 2005. Cellular DNA replicases: components and dynamics at the replication
fork. Annu. Rev. Biochem. 74:283–315
65. Kamada K. 2012. The GINS complex: structure and function. Subcell. Biochem. 62:135–56
66. Kasiviswanathan R, Shin J-H, Kelman Z. 2005. Interactions between the archaeal Cdc6 and MCM
proteins modulate their biochemical properties. Nucleic Acids Res. 33:4940–50
67. KasiviswanathanR, Shin J-H, Melamud E, Kelman Z. 2004. Biochemical characterization of the Methano-
thermobacter thermautotrophicus minichromosome maintenance (MCM) helicase N-terminal domains.
J. Biol. Chem. 279:28358–66
68. Kelch BA, Makino DL, O’Donnell M, Kuriyan J. 2011. How a DNA polymerase clamp loader opens a
sliding clamp. Science 334:1675–80
69. Kelch BA, Makino DL, O’Donnell M, Kuriyan J. 2012. Clamp loader ATPases and the evolution of
DNA replication machinery. BMC Biol. 10:34
92 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
70. Kelly TJ, Simancek P, Brush GS. 1998. Identification and characterization of a single-stranded DNA-
binding protein from the archaeon Methanococcus jannaschii.Proc. Natl. Acad. Sci. USA 95:14634–39
71. Kelman LM, Kelman Z. 2003. Archaea: an archetype for replication initiation studies? Mol. Microbiol.
48:605–15
72. Kelman LM, Kelman Z. 2004. Multiple origins of replication in archaea. Trends Microbiol. 12:399–401
73. Kelman Z. 2000. DNA replication in the third domain (of life). Curr. Protein Pept. Sci. 1:139–54
74. Kelman Z. 2000. The replication origin of archaea is finally revealed. Trends Biochem. Sci. 25:521–23
75. Kelman Z, Finkelstein J, O’Donnell M. 1995. Why have six-fold symmetry? Curr. Biol. 5:1239–42
76. Kelman Z, Hurwitz J. 1998. Protein-PCNA interactions: a DNA-scanning mechanism? Trends Biochem.
Sci. 23:236–38
77. Kelman Z, Hurwitz J. 2000. A unique organization of the protein subunits of the DNA polymerase clamp
loader in the archaeon Methanobacterium thermoautotrophicum H. J. Biol. Chem. 275:7327–36
78. Kelman Z, Hurwitz J. 2003. Structural lessons in DNA replication from the third domain of life. Nat.
Struct. Biol. 10:148–50
79. Kelman Z, Lee JK, Hurwitz J. 1999. The single minichromosome maintenance protein of Methanobac-
terium thermoautotrophicum H contains DNA helicase activity. Proc. Natl. Acad. Sci. USA 96:14783–88
80. Kelman Z, O’Donnell M. 1995. Structural and functional similarities of prokaryotic and eukaryotic DNA
polymerase sliding clamps. Nucleic Acids Res. 23:3613–20
81. Kelman Z, Pietrokovski S, Hurwitz J. 1999. Isolation and characterization of a split B-type DNA poly-
merase from the archaeon Methanobacterium thermoautotrophicum H. J. Biol. Chem. 274:28751–61
82. Kelman Z, Yuzhakov A, Andjelkovic J, O’Donnell M. 1998. Devoted to the lagging strand: the χsubunit
of DNA polymerase III holoenzyme contacts SSB to promote processive elongation and sliding clamp
assembly. EMBO J. 17:2436–49
83. Kennelly PJ. 2003. Archaeal protein kinases and protein phosphatases: insights from genomics and
biochemistry. Biochem. J. 370:373–89
84. Kerr ID, Wadsworth RI, Cubeddu L, Blankenfeldt W, Naismith JH, White MF. 2003. Insights into
ssDNA recognition by the OB fold from a structural and thermodynamic study of Sulfolobus SSB protein.
EMBO J. 22:2561–70
85. Kesti T, Flick K, Keranen S, Syvaoja JE, Wittenberg C. 1999. DNA polymerase catalytic domains are
dispensable for DNA replication, DNA repair, and cell viability. Mol. Cell 3:679–85
86. Kletzin A. 1992. Molecular characterisation of a DNA ligase gene of the extremely thermophilic archaeon
Desulfurolobus ambivalens shows close phylogenetic relationship to eukaryotic ligases. Nucleic Acids Res.
20:5389–96
87. Klinge S, Hirst J, Maman JD, Krude T, Pellegrini L. 2007. An iron-sulfur domain of the eukaryotic
primase is essential for RNA primer synthesis. Nat. Struct. Mol. Biol. 14:875–77
88. Komori K, Ishino Y. 2001. Replication protein A in Pyrococcus furiosus is involved in homologous DNA
recombination. J. Biol. Chem. 276:25654–60
89. Kornberg A, Baker TA. 1992. DNA Replication. New York: W.H. Freeman. 931 pp.
90. Krastanova I, Sannino V, Amenitsch H, Gileadi O, Pisani FM, Onesti S. 2012. Structural and functional
insights into the DNA replication factor Cdc45 reveal an evolutionary relationship to the DHH family
of phosphoesterases. J. Biol. Chem. 287:4121–28
91. Krueger S, Shin J-H, Curtis JE, Rubinson KA, Kelman Z. 2014. The solution structure of full-length
dodecameric MCM by SANS and molecular modeling. Proteins. In press
92. Krueger S, Shin J-H, Raghunandan S, Curtis JE, Kelman Z. 2011. Atomistic ensemble modeling and
small-angle neutron scattering of intrinsically disordered protein complexes: applied to minichromosome
maintenance protein. Biophys. J. 101:2999–3007
93. Krupovic M, Gribaldo S, Bamford DH, Forterre P. 2010. The evolutionary history of archaeal MCM
helicases: a case study of vertical evolution combined with hitch-hiking of mobile genetic elements. Mol.
Biol. Evol. 27:2716–32
94. Langston LD, Indiani C, O’Donnell M. 2009. Whither the replisome: emerging perspectives on the
dynamic nature of the DNA replication machinery. Cell Cycle 8:2686–91
95. Lao-Sirieix SH, Nookala RK, Roversi P, Bell SD, Pellegrini L. 2005. Structure of the heterodimeric
core primase. Nat. Struct. Mol. Biol. 12:1137–44
www.annualreviews.org Archaeal DNA Replication 93
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
96. Lao-Sirieix SH, Pellegrini L, Bell SD. 2005. The promiscuous primase. Trends Genet. 21:568–72
97. Lee C, Hong B, Choi JM, Kim Y, Watanabe S, et al. 2004. Structural basis for inhibition of the replication
licensing factor Cdt1 by geminin. Nature 430:913–17
98. Leonard AC, Grimwade JE. 2011. Regulation of DnaA assembly and activity: taking directions from the
genome. Annu. Rev. Microbiol. 65:19–35
99. Li Z, Huang RY, Yopp DC, Hileman TH, Santangelo TJ, et al. 2014. A novel mechanism for regulating
the activity of proliferating cell nuclear antigen by a small protein. Nucleic Acids Res. 42:5776–89
100. Li Z, Kelman LM, Kelman Z. 2013. Thermococcus kodakarensis DNA replication. Biochem. Soc. Trans.
41:332–38
101. Li Z, Pan M, Santangelo TJ, Chemnitz W, Yuan W, et al. 2011. A novel DNA nuclease is stimulated
by association with the GINS complex. Nucleic Acids Res. 39:6114–23
102. Li Z, Santangelo TJ, ˇ
Cubo ˇ
nov´
a L, Reeve JN, Kelman Z. 2010. Affinity purification of an archaeal DNA
replication protein network. MBio 1:e00221–10
103. Lindas AC, Bernander R. 2013. The cell cycle of archaea. Nat. Rev. Microbiol. 11:627–38
104. Liu W, Pucci B, Rossi M, Pisani FM, Ladenstein R. 2008. Structural analysis of the Sulfolobus solfataricus
MCM protein N-terminal domain. Nucleic Acids Res. 36:3235–43
105. Lopez P, Philippe H, Myllykallio H, Forterre P. 1999. Identification of putative chromosomal origins
of replication in Archaea. Mol. Microbiol. 32:883–86
106. Lundgren M, Andersson A, Chen L, Nilsson P, Bernander R. 2004. Three replication origins in Sulfolobus
species: synchronous initiation of chromosome replication and asynchronous termination. Proc. Natl.
Acad. Sci. USA 101:7046–51
107. Lundgren M, Bernander R. 2007. Genome-wide transcription map of an archaeal cell cycle. Proc. Natl.
Acad. Sci. USA 104:2939–44
108. MacNeill SA. 2010. Structure and function of the GINS complex, a key component of the eukaryotic
replisome. Biochem. J. 425:489–500
109. MacNeill SA. 2011. Protein-protein interactions in the archaeal core replisome. Biochem. Soc. Trans.
39:163–68
110. Makarova KS, Koonin EV. 2013. Archaeology of eukaryotic DNA replication. Cold Spring Harb. Perspect.
Biol. 5:a012963
111. Makarova KS, Koonin EV, Kelman Z. 2012. The CMG (CDC45/RecJ, MCM, GINS) complex is a
conserved component of the DNA replication system in all archaea and eukaryotes. Biol. Direct 7:7
112. Makarova KS, Wolf YI, Mekhedov SL, Mirkin BG, Koonin EV. 2005. Ancestral paralogs and pseu-
doparalogs and their role in the emergence of the eukaryotic cell. Nucleic Acids Res. 33:4626–38
113. Makarova KS, Yutin N, Bell SD, Koonin EV. 2010. Evolution of diverse cell division and vesicle forma-
tion systems in Archaea. Nat. Rev. Microbiol. 8:731–41
114. Marinsek N, Barry ER, Makarova KS, Dionne I, Koonin EV, Bell SD. 2006. GINS, a central nexus in
the archaeal DNA replication fork. EMBO Rep. 7:539–45
115. Masai H, Matsumoto S, You Z, Yoshizawa-Sugata N, Oda M. 2010. Eukaryotic chromosome DNA
replication: where, when, and how? Annu. Rev. Biochem. 79:89–130
116. Matsui E, Nishio M, Yokoyama H, Harata K, Darnis S, Matsui I. 2003. Distinct domain functions
regulating de novo DNA synthesis of thermostable DNA primase from hyperthermophile Pyrococcus
horikoshii.Biochemistry 42:14968–76
117. Matsui I, Urushibata Y, Shen Y, Matsui E, Yokoyama H. 2011. Novel structure of an N-terminal domain
that is crucial for the dimeric assembly and DNA-binding of an archaeal DNA polymerase D large subunit
from Pyrococcus horikoshii.FEBS Lett. 585:452–58
118. Matsunaga F, Forterre P, Ishino Y, Myllykallio H. 2001. In vivo interactions of archaeal Cdc6/Orc1 and
minichromosome maintenance protein with the replication origin. Proc. Natl. Acad. Sci. USA 98:11152–
57
119. Matsunaga F, Norais C, Forterre P, Myllykallio H. 2003. Identification of short “eukaryotic” Okazaki
fragments synthesized from a prokaryotic replication origin. EMBO Rep. 4:154–58
120. Maupin-Furlow JA. 2013. Ubiquitin-like proteins and their roles in archaea. Trends Microbiol. 21:31–38
94 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
121. Mayanagi K, Kiyonari S, Saito M, Shirai T, Ishino Y, Morikawa K. 2009. Mechanism of replication
machinery assembly as revealed by the DNA ligase-PCNA-DNA complex architecture. Proc. Natl. Acad.
Sci. USA 106:4647–52
122. McGeoch AT, Trakselis MA, Laskey RA, Bell SD. 2005. Organization of the archaeal MCM complex
on DNA and implications for the helicase mechanism. Nat. Struct. Mol. Biol. 12:756–62
123. Medagli B, Onesti S. 2013. Structure and mechanism of hexameric helicases. Adv. Exp. Med. Biol. 767:75–
95
124. Miyata T, Oyama T, Mayanagi K, Ishino S, Ishino Y, Morikawa K. 2004. The clamp-loading complex
for processive DNA replication. Nat. Struct. Mol. Biol. 11:632–36
125. Morgunova E, Gray FC, MacNeill SA, Ladenstein R. 2009. Structural insights into the adaptation
of proliferating cell nuclear antigen (PCNA) from Haloferax volcanii to a high-salt environment. Acta
Crystallogr. D 65:1081–88
126. Myllykallio H, Lopez P, Lopez-Garcia P, Heilig R, Saurin W, et al. 2000. Bacterial mode of replication
with eukaryotic-like machinery in a hyperthermophilic archaeon. Science 288:2212–15
127. O’Donnell M, Langston L, Stillman B. 2013. Principles and concepts of DNA replication in bacteria,
archaea, and eukarya. Cold Spring Harb. Perspect. Biol. 5:a010108
128. Ogino H, Ishino S, Mayanagi K, Haugland GT, Birkeland NK, et al. 2011. The GINS complex from the
thermophilic archaeon, Thermoplasma acidophilum may function as a homotetramer in DNA replication.
Extremophiles 15:529–39
129. Onesti S, MacNeill SA. 2013. Structure and evolutionary origins of the CMG complex. Chromosoma
122:47–53
130. Oyama T, Ishino S, Fujino S, Ogino H, Shirai T, et al. 2011. Architectures of archaeal GINS complexes,
essential DNA replication initiation factors. BMC Biol. 9:28
131. Pan M, Kelman LM, Kelman Z. 2011. The archaeal PCNA proteins. Biochem. Soc. Trans. 39:20–24
132. Pan M, Santangelo TJ, ˇ
Cubo ˇ
nov´
a L, Li Z, Metangmo H, et al. 2013. Thermococcus kodakarensis has two
functional PCNA homologues but only one is required for viability. Extremophiles 17:453–61
133. Pan M, Santangelo TJ, Li Z, Reeve JN, Kelman Z. 2011. Thermococcus kodakarensis encodes three MCM
homologs but only one is essential. Nucleic Acids Res. 39:9671–80
134. Pascal JM, O’Brien PJ, Tomkinson AE, Ellenberger T. 2004. Human DNA ligase I completely encircles
and partially unwinds nicked DNA. Nature 432:473–78
135. Pascal JM, Tsodikov OV, Hura GL, Song W, Cotner EA, et al. 2006. A flexible interface between DNA
ligase and PCNA supports conformational switching and efficient ligation of DNA. Mol. Cell 24:279–91
136. Paytubi S, McMahon SA, Graham S, Liu H, Botting CH, et al. 2012. Displacement of the canonical
single-stranded DNA-binding protein in the Thermoproteales. Proc. Natl. Acad. Sci. USA 109:E398–405
137. Pellegrini L. 2012. The Pol α-primase complex. Subcell. Biochem. 62:157–69
138. Pelve EA, Lindas AC, Knoppel A, Mira A, Bernander R. 2012. Four chromosome replication origins in
the archaeon Pyrobaculum calidifontis.Mol. Microbiol. 85:986–95
139. Perler FB, Kumar S, Kong H. 1996. Thermostable DNA polymerases. Adv. Protein Chem. 48:377–435
140. Pluchon PF, Fouqueau T, Creze C, Laurent S, Briffotaux J, et al. 2013. An extended network of genomic
maintenance in the archaeon Pyrococcus abyssi highlights unexpected associations between eucaryotic
homologs. PLoS ONE 8:e79707
141. Poplawski A, Grabowski B, Long SE, Kelman Z. 2001. The zinc finger domain of the archaeal minichro-
mosome maintenance protein is required for helicase activity. J. Biol. Chem. 276:49371–77
142. Prakash A, Borgstahl GE. 2012. The structure and function of replication protein A in DNA replication.
Subcell. Biochem. 62:171–96
143. Raymann K, Forterre P, Brochier-Armanet C, Gribaldo S. 2014. Global phylogenomic analysis disen-
tangles the complex evolutionary history of DNA replication in Archaea. Genome Biol. Evol. 6:192–212
144. Reimann J, Esser D, Orell A, Amman F, Pham TK, et al. 2013. Archaeal signal transduction: impact of
protein phosphatase deletions on cell size, motility and energy metabolism in Sulfolobus acidocaldarius.
Mol. Cell. Proteomics 12:3908–23
145. Remus D, Beuron F, Tolun G, Griffith JD, Morris EP, Diffley JF. 2009. Concerted loading of Mcm2-7
double hexamers around DNA during DNA replication origin licensing. Cell 139:719–30
www.annualreviews.org Archaeal DNA Replication 95
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
146. Reyes-Lamothe R, Nicolas E, Sherratt DJ. 2012. Chromosome replication and segregation in bacteria.
Annu. Rev. Genet. 46:121–43
147. Robbins JB, McKinney MC, Guzman CE, Sriratana B, Fitz-Gibbon S, et al. 2005. The euryarchaeota,
nature’s medium for engineering of single-stranded DNA-binding proteins. J. Biol. Chem. 280:15325–39
148. Robinson NP, Bell SD. 2007. Extrachromosomal element capture and the evolution of multiple repli-
cation origins in archaeal chromosomes. Proc. Natl. Acad. Sci. USA 104:5806–11
149. Robinson NP, Dionne I, Lundgren M, Marsh VL, Bernander R, Bell SD. 2004. Identification of two
origins of replication in the single chromosome of the archaeon Sulfolobus solfataricus.Cell 116:25–38
150. Sakakibara N, Kasiviswanathan R, Melamud E, Han M, Schwarz FP, Kelman Z. 2008. Coupling of
DNA binding and helicase activity is mediated by a conserved loop in the MCM protein. Nucleic Acids
Res. 36:1309–20
151. Sakakibara N, Kelman LM, Kelman Z. 2009. How is the archaeal MCM helicase assembled at the origin?
Possible mechanisms. Biochem. Soc. Trans. 37:7–11
152. Sakakibara N, Kelman LM, Kelman Z. 2009. Unwinding the structure and function of the archaeal
MCM helicase. Mol. Microbiol. 72:286–96
153. Samson RY, Bell SD. 2011. Cell cycles and cell division in the archaea. Curr. Opin. Microbiol. 14:350–56
154. Samson RY, Xu Y, Gadelha C, Stone TA, Faqiri JN, et al. 2013. Specificity and function of archaeal
DNA replication initiator proteins. Cell Rep. 3:485–96
155. Sarmiento F, Mrazek J, Whitman WB. 2013. Genome-scale analysis of gene function in the hy-
drogenotrophic methanogenic archaeon Methanococcus maripaludis.Proc. Natl. Acad. Sci. USA 110:4726–
31
156. Shechter DF, Ying CY, Gautier J. 2000. The intrinsic DNA helicase activity of Methanobacterium ther-
moautotrophicum H minichromosome maintenance protein. J. Biol. Chem. 275:15049–59
157. Shen Y, Tang XF, Matsui E, Matsui I. 2004. Subunit interaction and regulation of activity through
terminal domains of the family D DNA polymerase from Pyrococcus horikoshii.Biochem. Soc. Trans. 32:245–
49
158. Shin J-H, Grabowski B, Kasiviswanathan R, Bell SD, Kelman Z. 2003. Regulation of minichromosome
maintenance helicase activity by Cdc6. J. Biol. Chem. 278:38059–67
159. Shin J-H, Heo GY, Kelman Z. 2008. The Methanothermobacter thermautotrophicus Cdc6-2 protein, the
putative helicase loader, dissociates the minichromosome maintenance helicase. J. Bacteriol. 190:4091–94
160. Shin J-H, Heo G-Y, Kelman Z. 2009. The Methanothermobacter thermautotrophicus MCM helicase is
active as a hexameric ring. J. Biol. Chem. 284:540–46
161. Shin J-H, Kelman Z. 2006. The replicative helicases of bacteria, archaea and eukarya can unwind RNA-
DNA hybrid substrates. J. Biol. Chem. 281:26914–21
162. Shin J-H, Santangelo TJ, Xie Y, Reeve JN, Kelman Z. 2007. Archaeal minichromosome maintenance
(MCM) helicase can unwind DNA bound by archaeal histones and transcription factors. J. Biol. Chem.
282:4908–15
163. Skowyra A, MacNeill SA. 2012. Identification of essential and non-essential single-stranded DNA-
binding proteins in a model archaeal organism. Nucleic Acids Res. 40:1077–90
164. Slaymaker IM, Chen XS. 2012. MCM structure and mechanics: what we have learned from archaeal
MCM. Subcell. Biochem. 62:89–111
165. Sriskanda V, Kelman Z, Hurwitz J, Shuman S. 2000. Characterization of an ATP-dependent DNA ligase
from the thermophilic archaeon Methanobacterium thermoautotrophicum.Nucleic Acids Res. 28:2221–28
166. Stelter M, Gutsche I, Kapp U, Bazin A, Bajic G, et al. 2012. Architecture of a dodecameric bacterial
replicative helicase. Structure 20:554–64
167. Swiatek A, MacNeill SA. 2010. The archaeo-eukaryotic GINS proteins and the archaeal p rimase catalytic
subunit PriS share a common domain. Biol. Direct 5:17
168. Takahashi TS, Wigley DB, Walter JC. 2005. Pumps, paradoxes and ploughshares: mechanism of the
MCM2-7 DNA helicase. Trends Biochem. Sci. 30:437–44
169. Uemori T, Sato Y, Kato I, Doi H, Ishino Y. 1997. A novel DNA polymerase in the hyperthermophilic
archaeon, Pyrococcus furiosus: gene cloning, expression, and characterization. Genes Cells 2:499–512
170. Vivona JB, Kelman Z. 2003. The diverse spectrum of sliding clamp interacting proteins. FEBS Lett.
546:167–72
96 Kelman ·Kelman
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48CH04-Kelman ARI 13 October 2014 8:44
171. Wadsworth RI, White MF. 2001. Identification and properties of the crenarchaeal single-stranded DNA
binding protein from Sulfolobus solfataricus.Nucleic Acids Res. 29:914–20
172. Wahle E, Lasken RS, Kornberg A. 1989. The dnaB-dnaC replication protein complex of Escherichia coli.
II. Role of the complex in mobilizing dnaB functions. J. Biol. Chem. 264:2469–75
173. Warbrick E, Heatherington W, Lane DP, Glover DM. 1998. PCNA binding proteins in Drosophila
melanogaster: the analysis of a conserved PCNA binding domain. Nucleic Acids Res. 26:3925–32
174. Wigley DB. 2009. ORC proteins: marking the start. Curr. Opin. Struct. Biol. 19:72–78
175. Williams GJ, Johnson K, Rudolf J, McMahon SA, Carter L, et al. 2006. Structure of the heterotrimeric
PCNA from Sulfolobus solfataricus.Acta Crystallogr. Sect. F 62:944–48
176. Winter JA, Christofi P, Morroll S, Bunting KA. 2009. The crystal structure of Haloferax volcanii pro-
liferating cell nuclear antigen reveals unique surface charge characteristics due to halophilic adaptation.
BMC Struct. Biol. 9:55
177. Woese CR, Fox GE. 1977. Phylogenetic structure of the prokaryotic domain: the primary kingdoms.
Proc. Natl. Acad. Sci. USA 74:5088–90
178. Wu K, Lai X, Guo X, Hu J, Xiang X, Huang L. 2007. Interplay between primase and replication factor
C in the hyperthermophilic archaeon Sulfolobus solfataricus.Mol. Microbiol. 63:826–37
179. Yamasaki K, Urushibata Y, Yamasaki T, Arisaka F, Matsui I. 2010. Solution structure of the N-terminal
domain of the archaeal D-family DNA polymerase small subunit reveals evolutionary relationship to
eukaryotic B-family polymerases. FEBS Lett. 584:3370–75
180. Yao N, Turner J, Kelman Z, Stukenberg PT, Dean F, et al. 1996. Clamp loading, unloading and intrinsic
stability of the PCNA, βand gp45 sliding clamps of human, E. coli and T4 replicases. Genes Cells 1:101–13
181. Yao NY, O’Donnell M. 2012. The RFC clamp loader: structure and function. Subcell. Biochem. 62:259–79
182. Yuzhakov A, Kelman Z, Hurwitz J, O’Donnell M. 1999. Multiple competition reactions for RPA order
the assembly of the DNA polymerase δholoenzyme. EMBO J. 18:6189–99
183. Yuzhakov A, Kelman Z, O’Donnell M. 1999. Trading places on DNA: a three-point switch underlies
primer handoff from primase to the replicative DNA polymerase. Cell 96:153–63
184. Zhang R, Zhang CT. 2003. Multiple replication origins of the archaeon Halobacterium species NRC-1.
Biochem. Biophys.Res. Commun. 302:728–34
185. Zhao A, Gray FC, MacNeill SA. 2006. ATP- and NAD+-dependent DNA ligases share an essential
function in the halophilic archaeon Haloferax volcanii.Mol. Microbiol. 59:743–52
www.annualreviews.org Archaeal DNA Replication 97
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48-FrontMatter ARI 22 October 2014 14:49
Annual Review of
Genetics
Volume 48, 2014
Contents
L1 Retrotransposons and Somatic Mosaicism in the Brain
Sandra R. Richardson, Santiago Morell, and Geoffrey J. Faulkner ppppppppppppppppppppppppp1
Factors Underlying Restricted Crossover Localization
in Barley Meiosis
James D. Higgins, Kim Osman, Gareth H. Jones, and F. Chris H. Franklin pppppppppppp29
pENCODE: A Plant Encyclopedia of DNA Elements
Amanda K. Lane, Chad E. Niederhuth, Lexiang Ji, and Robert J. Schmitz pppppppppppppp49
Archaeal DNA Replication
Lori M. Kelman and Zvi Kelman pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp71
Molecular Genetic Dissection of Quantitative Trait Loci Regulating
Rice Grain Size
Jianru Zuo and Jiayang Li ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp99
Exploring Developmental and Physiological Functions of Fatty Acid
and Lipid Variants Through Worm and Fly Genetics
Huanhu Zhu and Min Han pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp119
Quality Control and Infiltration of Translation by Amino Acids
Outside of the Genetic Code
Tammy Bullwinkle, Beth Lazazzera, and Michael Ibba pppppppppppppppppppppppppppppppppp149
Vulnerabilities on the Lagging-Strand Template: Opportunities for
Mobile Elements
Ashwana D. Fricker and Joseph E. Peters ppppppppppppppppppppppppppppppppppppppppppppppppp167
Self-Organization of Meiotic Recombination Initiation: General
Principles and Molecular Pathways
Scott Keeney, Julian Lange, and Neeman Mohibullah pppppppppppppppppppppppppppppppppppp187
Cancer: Evolution Within a Lifetime
Marco Gerlinger, Nicholas McGranahan, Sally M. Dewhurst, Rebecca A. Burrell,
Ian Tomlinson, and Charles Swanton ppppppppppppppppppppppppppppppppppppppppppppppppppp215
Diverse Epigenetic Mechanisms of Human Disease
Emily Brookes and Yang Shi pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp237
From Egg to Gastrula: How the Cell Cycle Is Remodeled During the
Drosophila Mid-Blastula Transition
Jeffrey A. Farrell and Patrick H. O’Farrell ppppppppppppppppppppppppppppppppppppppppppppppp269
v
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.
GE48-FrontMatter ARI 22 October 2014 14:49
Cellular and Molecular Mechanisms of Single and Collective Cell
Migrations in Drosophila: Themes and Variations
Shirin M. Pocha and Denise J. Montell pppppppppppppppppppppppppppppppppppppppppppppppppppp295
The Structure and Regulation of Flagella in Bacillus subtilis
Sampriti Mukherjee and Daniel B. Kearns pppppppppppppppppppppppppppppppppppppppppppppppp319
Transcription-Associated Mutagenesis
Sue Jinks-Robertson and Ashok S. Bhagwat ppppppppppppppppppppppppppppppppppppppppppppppp341
Gastrointestinal Microbiota–Mediated Control of Enteric Pathogens
Sophie Yurist-Doutsch, Marie-Claire Arrieta, Stefanie L. Vogt, and B. Brett Finlay pp361
The Relations Between Recombination Rate and Patterns of Molecular
Variation and Evolution in Drosophila
Brian Charlesworth and Jos´eL.Campos ppppppppppppppppppppppppppppppppppppppppppppppppppp383
The Genetics of Neisseria Species
Ella Rotman and H. Steven Seifert pppppppppppppppppppppppppppppppppppppppppppppppppppppppp405
Regulation of Transcription by Long Noncoding RNAs
Roberto Bonasio and Ramin Shiekhattar ppppppppppppppppppppppppppppppppppppppppppppppppppp433
Centromeric Heterochromatin: The Primordial Segregation
Machine
Kerry S. Bloom pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp457
Nonadditive Gene Expression in Polyploids
Mi-Jeong Yoo, Xiaoxian Liu, J. Chris Pires, Pamela S. Soltis,
and Douglas E. Soltis ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp485
Lineage Sorting in Apes
Thomas Mailund, Kasper Munch, and Mikkel Heide Schierup pppppppppppppppppppppppppp519
Messenger RNA Degradation in Bacterial Cells
Monica P. Hui, Patricia L. Foley, and Joel G. Belasco pppppppppppppppppppppppppppppppppppp537
Population Genomics of Transposable Elements in Drosophila
Maite G. Barr´on, Anna-Sophie Fiston-Lavier, Dmitri A. Petrov,
and Josefa Gonz´alez pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp561
Genetic, Epigenetic, and Environmental Contributions to Neural
Tube Closure
Jonathan J. Wilde, Juliette R. Petersen, and Lee Niswander ppppppppppppppppppppppppppppp583
Errata
An online log of corrections to Annual Review of Genetics articles may be found at
http://www.annualreviews.org/errata/genet
vi Contents
Annu. Rev. Genet. 2014.48:71-97. Downloaded from www.annualreviews.org
Access provided by University of Maryland - College Park on 11/25/14. For personal use only.

File (1)

ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The origin of the eukaryotic cell can be regarded as one of the hallmarks in the history of life on our planet. The apparent genomic chimerism in eukaryotic genomes is currently best explained by invoking a cellular fusion at the root of the eukaryotes that involves one archaeal and one or more bacterial components. Here, we use a phylogenomics approach to reevaluate the evolutionary affiliation between Archaea and eukaryotes, and provide further support for scenarios in which the nuclear lineage in eukaryotes emerged from within the archaeal radiation, displaying a strong phylogenetic affiliation with, or even within, the archaeal TACK superphylum. Further taxonomic sampling of archaeal genomes in this superphylum will certainly provide a better resolution in the events that have been instrumental for the emergence of the eukaryotic lineage.
Article
Full-text available
Accurate DNA replication is essential for maintenance of every genome. All archaeal genomes except Crenarchaea, encode for a member of Family B (polB) and Family D (polD) DNA polymerases. Gene deletion studies in Thermococcus kodakaraensis and Methanococcus maripaludis show that polD is the only essential DNA polymerase in these organisms. Thus, polD may be the primary replicative DNA polymerase for both leading and lagging strand synthesis. To understand this unique archaeal enzyme, we report the biochemical characterization of a heterodimeric polD from Thermococcus. PolD contains both DNA polymerase and proofreading 3′–5′ exonuclease activities to ensure efficient and accurate genome duplication. The polD incorporation fidelity was determined for the first time. Despite containing 3′–5′ exonuclease proofreading activity, polD has a relatively high error rate (95 × 10−5) compared to polB (19 × 10−5) and at least 10-fold higher than the polB DNA polymerases from yeast (polε and polδ) or Escherichia coli DNA polIII holoenzyme. The implications of polD fidelity and biochemical properties in leading and lagging strand synthesis are discussed. Electronic supplementary material The online version of this article (doi:10.1007/s00792-014-0646-9) contains supplementary material, which is available to authorized users.
Article
Full-text available
Living organisms are divided into three domains: Bacteria, Archaea, and Eukarya. Whereas Bacteria and Archaea are both prokaryotes, proteins involved in information processes; replication, transcription, and translation, are more similar in Archaea and Eukarya. Here the history of the research on archaeal DNA replication is summarized and the future of the field is discussed.
Article
Full-text available
The archaeal machinery responsible for DNA replication is largely homologous to that of eukaryotes and is clearly distinct from its bacterial counterpart. Moreover, it shows high diversity in the various archaeal lineages, including different sets of components, heterogeneous taxonomic distribution, and a large number of additional copies that are sometimes highly divergent. This has made the evolutionary history of this cellular system particularly challenging to dissect. Here, we have carried out an exhaustive identification of homologs of all major replication components in over 140 complete archaeal genomes. Phylogenomic analysis allowed assigning them to either a conserved and probably essential 'core' of replication components that were mainly vertically inherited, or to a variable and highly divergent 'shell' of extra copies that have likely arisen from integrative elements. This suggests that replication proteins are frequently exchanged between extra-chromosomal elements and cellular genomes. Our study allowed clarifying the history that shaped this key cellular process (ancestral components, horizontal gene transfers, gene losses), providing important evolutionary and functional information. Finally, our precise identification of core components permitted to show that the phylogenetic signal carried by DNA replication is highly consistent with that harbored by two other key informational machineries (translation and transcription), strengthening the existence of a robust organismal tree for the Archaea.
Article
Full-text available
In Archaea, the proteins involved in the genetic information processing pathways, including DNA replication, transcription, and translation, share strong similarities with those of eukaryotes. Characterizations of components of the eukaryotic-type replication machinery complex provided many interesting insights into DNA replication in both domains. In contrast, DNA repair processes of hyperthermophilic archaea are less well understood and very little is known about the intertwining between DNA synthesis, repair and recombination pathways. The development of genetic system in hyperthermophilic archaea is still at a modest stage hampering the use of complementary approaches of reverse genetics and biochemistry to elucidate the function of new candidate DNA repair gene. To gain insights into genomic maintenance processes in hyperthermophilic archaea, a protein-interaction network centred on informational processes of Pyrococcus abyssi was generated by affinity purification coupled with mass spectrometry. The network consists of 132 interactions linking 87 proteins. These interactions give insights into the connections of DNA replication with recombination and repair, leading to the discovery of new archaeal components and of associations between eucaryotic homologs. Although this approach did not allow us to clearly delineate new DNA pathways, it provided numerous clues towards the function of new molecular complexes with the potential to better understand genomic maintenance processes in hyperthermophilic archaea. Among others, we found new potential partners of the replication clamp and demonstrated that the single strand DNA binding protein, Replication Protein A, enhances the transcription rate, in vitro, of RNA polymerase. This interaction map provides a valuable tool to explore new aspects of genome integrity in Archaea and also potentially in Eucaryotes.
Article
The discoveries of DNA supercoiling and DNA topoisomerases have been two of the most important breakthroughs in biology in the last century. Negative supercoiling is a critical feature of bacterial genomes and transient supercoiling is produced in all organisms by DNA binding proteins and/or various DNA tracking processes. DNA topoisomerases are major elements in cellular life and the plethora of natural antibiotics and antitumor drugs that target these enzymes testify for their importance. DNA topoisomerases I and II, catalyzing DNA strand transfer via single or double-strand breaks in DNA molecules respectively, originated and evolved to solve topological problems raised by the plectonemic coiling of two DNA strands in the double helix, and specialized topoisomerases acquired the ability to produce supercoiling. The first DNA topoisomerases were discovered in Escherichia coli (protein ω, DNA gyrase, Topo III, and Topo IV) and eukaryotic cells (Topo IB, Topo II, Topo III). Later on, new families and subfamilies of DNA topoisomerases were discovered in Archaea, the third domain of life (reverse gyrase, Topo V, Topo VI), challenging the prokaryote/eukaryote dichotomy. DNA topoisomerases are now classified into five families of homologous proteins (Topo IA, IB, IC, Topo IIA, IIB) based on structural similarities. These families have been divided into subfamilies, some of them characterized by unique enzymatic properties (gyrase, reverse gyrase). The distribution of these families and subfamilies do not overlap with the universal tree of life, and some subfamilies are specific for viruses. This suggests that viruses played an important role in the origin and distribution of DNA topoisomerases among cellular organisms.
Article
ENCODE projects exist for many eukaryotes, including humans, but as of yet no defined project exists for plants. A plant ENCODE would be invaluable to the research community and could be more readily produced than its metazoan equivalents by capitalizing on the preexisting infrastructure provided from similar projects. Collecting and normalizing plant epigenomic data for a range of species will facilitate hypothesis generation, cross-species comparisons, annotation of genomes, and an understanding of epigenomic functions throughout plant evolution. Here, we discuss the need for such a project, outline the challenges it faces, and suggest ways forward to build a plant ENCODE. Expected final online publication date for the Annual Review of Genetics Volume 48 is November 23, 2014. Please see http://www.annualreviews.org/catalog/pubdates.aspx for revised estimates.
Article
The solution structure of the full-length DNA helicase minichromosome maintenance (MCM) protein from Methanothermobacter thermautotrophicus was determined by small-angle neutron scattering (SANS) data together with all-atom molecular modeling. The data were fit best with a dodecamer (dimer of hexamers). The twelve monomers were linked together by the B/C domains, and the ATPase (AAA+) catalytic regions were found to be freely movable in the full-length dodecamer both in the presence and absence of Mg2+ and 50-meric ssDNA. In particular, the SANS data and molecular modeling indicate that all twelve AAA+ domains in the dodecamer lie approximately the same distance from the axis of the molecule, but the positions of the helix-turn-helix region at the C-terminus of each monomer differ. In addition, the A domain at the N-terminus of each monomer is tucked up next to the AAA+ domain for all twelve monomers of the dodecamer. Finally, binding of ssDNA does not lock the AAA+ domains in any specific position, which leaves them with the flexibility to move both for helicase function and for binding along the ssDNA. © Proteins 2014;. © 2014 Wiley Periodicals, Inc.
Article
Proliferating cell nuclear antigen (PCNA) forms a trimeric ring that associates with and influences the activity of many proteins participating in DNA metabolic processes and cell cycle progression. Previously, an uncharacterized small protein, encoded by TK0808 in the archaeon Thermococcus kodakarensis, was shown to stably interact with PCNA in vivo. Here, we show that this protein, designated Thermococcales inhibitor of PCNA (TIP), binds to PCNA in vitro and inhibits PCNA-dependent activities likely by preventing PCNA trimerization. Using hydrogen/deuterium exchange mass spectrometry and site-directed mutagenesis, the interacting regions of PCNA and TIP were identified. Most proteins bind to PCNA via a PCNA-interacting peptide (PIP) motif that interacts with the inter domain connecting loop (IDCL) on PCNA. TIP, however, lacks any known PCNA-interacting motif, suggesting a new mechanism for PCNA binding and regulation of PCNA-dependent activities, which may support the development of a new subclass of therapeutic biomolecules for inhibiting PCNA.

Linked Research