Download Structure and function of the GINS complex, a key component of the

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

DNA polymerase wikipedia , lookup

DNA damage theory of aging wikipedia , lookup

Genomics wikipedia , lookup

Deoxyribozyme wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

Non-coding DNA wikipedia , lookup

Cancer epigenetics wikipedia , lookup

Epigenetics of human development wikipedia , lookup

Gene wikipedia , lookup

Chromosome wikipedia , lookup

DNA supercoil wikipedia , lookup

Epigenomics wikipedia , lookup

Nucleosome wikipedia , lookup

Molecular cloning wikipedia , lookup

Protein moonlighting wikipedia , lookup

Primary transcript wikipedia , lookup

Microevolution wikipedia , lookup

Cre-Lox recombination wikipedia , lookup

DNA vaccination wikipedia , lookup

History of genetic engineering wikipedia , lookup

Extrachromosomal DNA wikipedia , lookup

DNA replication wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

Eukaryotic DNA replication wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Point mutation wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Helitron (biology) wikipedia , lookup

Replisome wikipedia , lookup

Polycomb Group Proteins and Cancer wikipedia , lookup

NEDD9 wikipedia , lookup

Transcript
Biochem. J. (2010) 425, 489–500 (Printed in Great Britain)
489
doi:10.1042/BJ20091531
REVIEW ARTICLE
Structure and function of the GINS complex, a key component of the
eukaryotic replisome
Stuart A. MACNEILL1
Biomedical Sciences Research Complex, School of Biology, University of St Andrews, North Haugh, St Andrews, Fife KY16 9ST, U.K.
High-fidelity chromosomal DNA replication is fundamental to
all forms of cellular life and requires the complex interplay of
a wide variety of essential and non-essential protein factors in a
spatially and temporally co-ordinated manner. In eukaryotes, the
GINS complex (from the Japanese go-ichi-ni-san meaning 5-12-3, after the four related subunits of the complex Sld5, Psf1,
Psf2 and Psf3) was recently identified as a novel factor essential
for both the initiation and elongation stages of the replication
process. Biochemical analysis has placed GINS at the heart
of the eukaryotic replication apparatus as a component of the
CMG [Cdc45–MCM (minichromosome maintenance) helicase–
GINS] complex that most likely serves as the replicative helicase,
unwinding duplex DNA ahead of the moving replication fork.
GINS homologues are found in the archaea and have been
shown to interact directly with the MCM helicase and with
primase, suggesting a central role for the complex in archaeal
chromosome replication also. The present review summarizes
current knowledge of the structure, function and evolution of
the GINS complex in eukaryotes and archaea, discusses possible
functions of the GINS complex and highlights recent results that
point to possible regulation of GINS function in response to DNA
damage.
INTRODUCTION
additional components of the replisome appear to assemble [6,7].
The function of the GINS complex has been analysed by a variety
of methods in a wide range of species and the three-dimensional
structure of the human GINS complex has been solved by X-ray
crystallography [8–10]. Archaeal GINS homologues have also
been identified [11–13], providing insights into GINS complex
evolution and core function, whereas evidence from yeast and
mammalian cells point to a possible secondary role for GINS
in chromosome segregation [14] and highlights the importance
of GINS in development and disease avoidance [15–19]. Finally,
recent results suggest that the GINS function might be regulated
in response to DNA damage [20].
Chromosomal DNA replication in all cells requires the complex
interplay of a large number of essential and non-essential protein
factors in a spatially and temporally co-ordinated manner [1].
In eukaryotic cells, replication is initiated at multiple sites
(replication origins) on each chromosome. Assembly of the
replication machinery begins during G1 , but multiple regulatory
mechanisms ensure that origin DNA unwinding and replication
initiation does not occur until S-phase and that replication occurs
once and only once in every cell cycle. As the origin DNA is
unwound, the replication machinery (replisome) is assembled and
the replication forks move bidirectionally from the origin into the
non-origin DNA [1,2].
Much of what is known of the events of replication initiation
has come through work on experimental organisms such
as the budding yeast Saccharomyces cerevisiae, the fission
yeast Schizosaccharomyces pombe and the African clawed
frog Xenopus laevis. Numerous factors essential for replication
initiation have been identified and their relationships to one
another deciphered in these model systems [1].
In the present review I summarize our current understanding of
the function, structure and evolution of the most recently identified
essential replication factor in eukaryotic cells, the tetrameric
GINS complex [3–5]. GINS (the name is an acronym for goichi-ni-san, the Japanese for 5-1-2-3, after the four subunits of the
complex Sld5, Psf1, Psf2 and Psf3) is essential for the initiation
and elongation stages of chromosome replication [3–5] and serves
as a key component of a protein complex that unwinds the duplex
DNA ahead of the moving replication fork and around which many
Key words: CMG complex, chromosome replication, DNA
helicase, DNA replication, GINS.
STEPWISE ASSEMBLY OF THE REPLISOME
The early events of chromosome replication are best understood
in the budding yeast, although it is highly likely that key features
of the initiation mechanism will be conserved across evolution
[1,2]. In budding yeast, replication origins are bound throughout
the cell cycle by a conserved six-subunit protein complex known
as ORC (origin recognition complex). During the G1 -phase of the
cell cycle, ORC is bound by the Cdc6 protein which then recruits
two additional factors, Cdt1 and the MCM (minichromosome
maintenance) helicase, to form the pre-RC (pre-replicative
complex) (shown in Figure 1). Assembly of the pre-RC is
tightly regulated, with protein phosphorylation in particular
playing an important role. Pre-RC assembly, also known as
replication licensing, occurs only during G1 , when levels of
the inhibitory CDK (cyclin-dependent kinase) are low, thereby
Abbreviations used: ATM, ataxia-telangiectasia mutated; ATR, ATM and Rad3 related; BiFC, bimolecular fluorescence complementation; CDK, cyclindependent kinase; ChIP, chromatin immunoprecipitation; CPC, chromosomal passenger complex; E, embryonic day; EM, electron microscopy; GFP,
green fluorescent protein; GINS, go-ichi-ni-san; HBD, hormone-binding domain; HDF, human dermal fibroblast; Hsp90, heat-shock protein of 90 kDa;
MCM, minichromosome maintenance; ORC, origin recognition complex; PCNA, proliferating-cell nuclear antigen; pre-LC, pre-loading complex; pre-RC,
pre-replicative complex; RecJdbh, RecJ DNA-binding domain homologue; RPC, replisome progression complex; siRNA, small interfering RNA; S-CDK,
S-phase CDK; ssDNA, single-stranded DNA; YFP, yellow fluorescent protein.
1
email [email protected]
c The Authors Journal compilation c 2010 Biochemical Society
490
Figure 1
S. A. MacNeill
Model for stepwise replisome assembly in budding yeast
ORC binds to replication origin DNA throughout the cell cycle. In G1 , Cdc6 binds to ORC and
recruits the hexameric MCM helicase and Cdt1 to form the pre-RC. Pre-RC assembly is said
to licence the origin for replication and can only occur in G1 , thereby preventing re-replication
of the same sequences in G2 . It is unclear whether MCM is loaded as a hexamer or a double
hexamer; for clarity, only a single hexamer is shown but at some point a second MCM hexamer
must be loaded to facilitate bidirectional replication. Cdc45–Sld3 appears to associate with
pre-RC-bound origins in G1 . The order and interdependence of the subsequent steps is unclear:
Sld3 is phosphorylated by S-CDK, apparently allowing it to bind to Dpb11 and recruit the
pre-LC, whose assembly is also CDK-dependent, to the origin. The pre-LC contains Dpb11,
Sld2 (S-CDK phosphorylation of which is required for binding to Dpb11), the leading strand
polymerase Pol ε and the GINS tetramer. Cdc45, MCM and GINS forms a tight complex (CMG)
that moves with the replication fork and which most likely functions as the replicative helicase.
Assembled around a single CMG complex is the multi-protein replisome progression complex
(RPC, indicated by the grey shading); a full list of RPC components can be found in the text. The
replisome is shown moving to the left (indicated by block arrow): Pol ε is shown synthesizing
the leading strand. See text for details and references.
and DDK substrates is key to understanding how replication
initiation is regulated. S-CDK substrates are abundant in yeast
but two essential proteins, Sld2 and Sld3, have been shown
to comprise the minimal set required for replication initiation
[21,22]. Phosphorylation of Sld3 by S-CDK on Thr600 and Ser622
is essential for cell viability, substituting these two residues with
non-phosphorylatable alanine residues blocks entry into S-phase.
The effect of the phosphorylation of Sld3 is to generate a binding
site from another essential replication factor, Dpb11 (equivalent
to TOPBP1 in human cells). The Dpb11 protein possesses four
BRCT domains that act in pairs to bind to phosphopeptides
[23] and it is the N-terminal pair that is responsible for binding
to phosphorylated Sld3 (Figure 1) [21,22]. The importance of
Thr600 and Ser622 phosphorylation in mediating Dpb11 binding is
shown by the fact that fusion of non-phosphorylatable Sld3 to
Dpb11 produces a protein that is able to rescue loss-of-function
mutations in both genes and partially bypass the requirement for
S-CDK activity to initiate S-phase [22]. Note that Sld3 also associates with another essential replication factor, Cdc45, which is
likely to be recruited to the pre-RC via its interactions with MCM
(Figure 1). Interestingly, a mutant form of Cdc45 (the Cdc45-JET
protein) can also partially bypass the requirement for S-CDK,
probably by promoting tighter Sld3–Dpb11 interactions [21].
The second essential S-CDK substrate in yeast is Sld2
(equivalent to RecQL4 in human cells). This protein is phosphorylated on multiple sites by S-CDK, but Thr84 phosphorylation
is crucial for Sld2 function [24]. The effect of Thr84
phosphorylation is also to create a binding site for Dpb11,
this time via the C-terminal Dpb11 BRCT domains (Figure 1).
By combining a mutant form of the Sld2 protein in which Thr84
is substituted with an aspartate residue to mimic phosphorylation
with the Sld3–Dpb11 fusion protein described above, it is
possible to completely bypass the requirement for S-CDK
in replication initiation [22]. Similar results were seen when
the phosphomimicking Sld2 mutant was combined with cells
expressing the Cdc45-JET protein and overproducing Dpb11 [21].
Dpb11, therefore, bridges phosphorylated Sld2 and Sld3, but
to what end? Although yet to be described in detail, Sld2 is
reportedly [25] part of a complex (termed the pre-loading complex
or pre-LC) which includes Dpb11, the leading strand polymerase
Polε and the subject of the present review, GINS, suggesting
perhaps that the function of Dpb11 in bridging Cdc45–Sld3 with
Sld2 is to recruit the pre-LC components to the origin to activate
the MCM helicase (Figure 1). The role of GINS in MCM helicase
function will be described in detail below.
Once the replisome is activated, the replication forks move
off bidirectionally from the origin DNA into the surrounding
sequences (Figure 1). ORC remains bound at the origin,
apparently intact, at least in yeast cells, whereas MCM, Cdc45,
GINS and Polε move with the fork [3,5,26–28]. The remainder
of the factors described above dissociate from the replisome
and in some cases are degraded or excluded from the nucleus.
Meanwhile, numerous additional protein factors – some essential,
some not – are recruited around the Cdc45–MCM–GINS core
[29] to form a larger macromolecular assembly known as the
RPC (replisome progression complex) [6,7,30]. This is discussed
further below.
IDENTIFICATION OF THE GINS COMPLEX
ensuring that replication occurs once, and only once, per
cell cycle [1].
Subsequent to pre-RC assembly, replication initiation is
promoted by the action of S-CDK (S-phase CDK) and a second
protein kinase, DDK (Cdc7-Dbf4) [1,2]. Identification of S-CDK
c The Authors Journal compilation c 2010 Biochemical Society
Eukaryotic GINS
Identification of the GINS complex was reported by two groups
working with budding yeast, but using very different experimental
approaches. As part of their ongoing genetic analysis of the role of
Structure and function of the GINS complex
Dpb11 in replication initiation, Araki and co-workers [31] isolated
mutations that were synthetically lethal in combination with the
temperature-sensitive dpb11-1 allele. These fell into a number
of distinct complementation groups, including SLD2 and SLD3
described above, and also SLD5 (SLD is an acronym for synthetic
lethal with dpb11-1). The SLD5 gene was cloned and shown
to encode a protein of 294 amino acids of unknown function. To
identify factors that might interact with SLD5, Takayama et al. [5]
first isolated temperature-sensitive alleles of SLD5, then screened
for multicopy suppressors of one of these, sld5-12, and in this
way isolated the PSF1 gene encoding a 208-amino-acid protein.
Next, a temperature-sensitive psf1 allele, psf1-1, was generated
and another new gene, PSF3, encoding a 194-amino-acid protein,
isolated as a multicopy suppressor. Although not isolated in
their screens, SLD5 was also shown to rescue psf1-1, albeit
weakly.
These genetic interactions suggested strongly that the products
of the budding yeast SLD5, PSF1 and PSF3 genes might
interact with each other. To test this, an epitope-tagged
form of the Psf1 protein was affinity-purified under native
conditions from exponentially growing yeast cells and was
shown to co-purify with Sld5 and Psf3, as well as with
a novel 213-amino-acid protein later designated Psf2 [5].
Overproduction of PSF2, like PSF1, was also shown to be able
to suppress to sld5-12. Further biochemical analysis showed
that all four proteins were stoichiometric components of a
100 kDa complex present at constant levels through the cell cycle;
the complex was named GINS.
While Takayama et al. [5] were treading a traditional yeast
molecular genetic path, Labib and co-workers [3] took an
altogether different approach. Previous work from a consortium
of laboratories [32] had identified a significant number of budding
yeast genes with essential functions but where comparative
protein sequence analysis gave no clues to their cellular roles.
In order to identify members of this group that were specifically
required for cell-cycle progression, temperature-sensitive mutant
forms of many of these essential proteins were constructed by
fusing a heat-inducible degron cassette [33] to their N-termini
[3]. At higher temperatures, the unfolded degron is recognized
by the Ubr1 protein, targeted for ubiquitylation and subsequently
degraded by the proteasome. Three of the four subunits of the
GINS complex were identified by this approach: Sld5, Psf1
and Psf2 (note that these were designated Cdc105, Cdc101 and
Cdc102 in the original paper describing this work) [3]. Purification
of Psf2 from yeast cell extracts demonstrated the physical
association of this protein with Sld5 and Psf1 and allowed the
subsequent identification of Psf3 [3].
In parallel with identification of budding yeast GINS, Takisawa
and colleagues [4] reported the isolation of the Xenopus GINS
proteins and showed that these too formed a heterotetrameric
GINS complex with a molecular mass of ∼ 100 kDa.
FUNCTION OF GINS
Budding yeast
Each of the four GINS proteins is essential for cell viability in
budding yeast. Inactivation of individual proteins by temperatureshifting cells carrying temperature-sensitive mutant alleles of
sld5 or psf1 [5] or heat-inducible degron alleles of sld5 or psf2
[3] results in cell-cycle arrest, with the arrested cells having a
single nucleus and displaying the large-budded (dumbbell) shape
that is typical of yeast replication mutants. Chromosomal DNA
replication is blocked under these conditions, particularly in cells
expressing the Sld5 and Psf2 degron fusion proteins [3].
491
To investigate GINS function in greater detail, ChIP
(chromatin immunoprecipitation) [34] was used to ask whether
the GINS proteins associated with replicating DNA in cells
proceeding slowly through S-phase at low temperature or at
low concentrations of the ribonucleotide reductase inhibitor HU
(hydroxyurea). In both cases, GINS proteins were found to
associate with origin sequences in early S-phase, before moving
with the replication fork from origin to non-origin DNA as
replication continued [5,28]. In addition to this, GINS was
shown by density substitution experiments to be required both
for replication fork establishment at origins and, importantly,
also for replication fork progression, demonstrating that the
GINS complex is not merely a passenger at the fork but an
active participant [3]. Consistent with GINS being required for
replication fork establishment at origins, neutral/neutral twodimensional gel analysis showed firing of the budding yeast
replication origin ARS1 to be reduced in sld5 and psf1 mutants
[5].
The origin association of GINS is dependent on both Sld3
and Dpb11 function: when sld5-3 or dpb11-26 cells are released
into S-phase from G1 at their restrictive temperatures, no binding
of GINS to origins can be detected by ChIP [5], suggesting
that this event is downstream of Sld3 and Dpb11 function (see
Figure 1). However, GINS function is required for Dpb11 to
associate with origins, indicating that GINS and Dpb11 origin
binding are mutually dependent events. As noted above, there are
indications that GINS and Dpb11 are present in a distinct complex,
termed the pre-LC, along with Sld2 and Polε [25]; it is tempting to
speculate that the observed interdependency of GINS and Dpb11
reflects the consequences of disruption of this complex.
GINS is also required for Cdc45 chromatin binding [5], but not
for initial recruitment of Cdc45 to origins during late G1 . However,
GINS is required to allow Cdc45 to move away from origins
as replication progresses, as well as for the stable association
of Cdc45 with origin sequences in cells in which replication is
stalled by inactivation of GINS using a heat-inducible degrontagged Psf2 protein [28].
Fission yeast
The four subunits of GINS in the fission yeast S. pombe are similar
in sequence to their budding yeast counterparts and are presumed
to form a similar heterotetrameric complex. As in budding yeast,
ChIP has been used to show that GINS associates with origin
DNA in early S-phase [35].
Several conditional lethal mutant forms of the fission yeast
GINS proteins have been used for analysis of S. pombe GINS
function, including traditional temperature-sensitive alleles of
psf2 [14,36] and psf3 [35], and hormone-dependent alleles of psf1
and psf2 [37,38]. The first of these to be isolated was the
temperature-sensitive psf2-209 allele, from a screen for mutations
that prevented re-replication of chromosomal DNA caused by an
inappropriate reduction of CDK activity [36]. When shifted to
the restrictive temperature, psf2-209 cells display a somewhat
heterogeneous phenotype: the cells are elongated, show some
nuclear abnormalities (discussed further below) and have DNA
contents in the 1C–2C range, suggesting that either entry into
S-phase or S-phase progression is delayed. Consistent with
the latter possibility, psf2-209 cells require the presence of
a functional DNA-damage checkpoint for viability. The cells
probably also display defects in pre-meiotic S-phase [36].
In contrast with psf2-209, psf3-1 was isolated in a targeted
screen for psf3 mutations. Cells carrying this temperaturesensitive allele undergo cell-cycle arrest when shifted to their
restrictive temperature, accumulating at early time points with
c The Authors Journal compilation c 2010 Biochemical Society
492
S. A. MacNeill
1C DNA content that is consistent with an early S-phase
block [35]. At later time points, cells with 2C DNA are seen,
indicating that some replication is possible in this mutant,
as well as cells with <1C DNA that are thought to be the
result of cell division occurring in the absence of complete
DNA replication. The sld3-1 mutation can be suppressed by
overproduction of Sld5, Psf1 and Psf2, and also by Sld2 [35],
and is synthetically lethal at the normal permissive temperature
with mutant alleles of cut5 and cdc20, encoding the fission
yeast Dpb11 homologue and the catalytic subunit of Polε
respectively, as well as with sld3 and cdc45. The GINS complex
appears also to be partially destabilized in psf3-1 cells; the
mutant Psf3-1 protein fails to co-purify with FLAG-tagged Psf2,
although pulldown of Sld5 and Psf1 is unaffected, indicating
that Psf3 is not required for formation of an Sld5–Psf1–Psf2
sub-complex.
As in budding yeast, binding of Cdc45 to origins (assayed by
ChIP) is abolished by GINS inactivation in psf3-1 cells. The same
is true of Cut5 and Drc1 (the fission yeast orthologues of Dpb11
and Sld2 respectively) and Polε, but not MCM and Sld3, both of
which associate normally with origin DNA. Conversely, both Cut5
and Sld3 are required for GINS binding to origins, but Cdc45 is
not [35].
Studies in several organisms have shown that fusing a
β-oestradiol HBD (hormone-binding domain) to a target protein
can result in the activity of the protein becoming regulated by
the hormone [39]. In certain cases, this appears to be caused
by Hsp90 (heat-shock protein of 90 kDa) binding to the HBD in
the absence of hormone and blocking protein–protein interactions
by steric hindrance. The inactivation seen is reversible, as
subsequent addition of β-oestradiol causes Hsp90 to dissociate
from the HDB. In fission yeast, fusing the β-oestradiol HBD to
Psf1 and Psf2 produces cells that require the constant presence
of β-oestradiol in the growth medium for cell viability [37,38].
Removal of β-oestradiol causes rapid cell-cycle arrest in early
S-phase and in the case of Psf2, nuclear delocalization of the
protein (Psf1 was not analysed in this way). The Psf1–HBD
mutant has been used to show that GINS function is required for
Cdc45 chromatin loading at S-phase initiation (as above), as well
as for maintenance of Cdc45 binding during the elongation stages
of replication [38]. GINS inactivation by β-oestradiol removal
also results in a failure to load Polε on to chromatin, as seen
previously [35]. Polα is loaded normally in these conditions, but
is not released from chromatin as S-phase cannot be completed in
the absence of GINS function.
Finally, interactions between GINS and MCM have been
investigated using BiFC (bimolecular fluorescence complementation) in fission yeast [40]. This technique relies on the
ability of the N- and C-terminal fragments of the YFP (yellow
fluorescent protein) to form a fluorescent complex when brought
together by the association of two interacting partners [41]. Using
BiFC, specific interactions between GINS and MCM mediated
by Psf1 and Mcm4 have been observed on chromatin during
S-phase, consistent with the role of the proteins at the replication
fork [40]. However, GINS–MCM interactions were also seen in
other phases of the cell cycle, perhaps due to the fact that once
they form a complex, the two fragments of YFP are extremely
reluctant to dissociate [41].
Xenopus
The frog Xenopus laevis has proved a very powerful model
for biochemical studies of eukaryotic chromosome replication.
Central to this has been the ability of extracts prepared from
c The Authors Journal compilation c 2010 Biochemical Society
Xenopus eggs to efficiently replicate exogenously added DNA
templates such as sperm chromatin or purified plasmid DNA and
to do so in a manner that preserves cell-cycle control of replication
licensing [42].
Xenopus GINS was the first higher eukaryotic GINS complex
to be functionally characterized [4]. The four subunits of frog
GINS are approx. 20–25 % identical (40–50 % similar) at the
amino acid level to the budding yeast proteins. Antibodies raised
against Sld5 co-immunoprecipitate Psf1, Psf2 and Psf3, indicating
that the four proteins co-exist in a complex. Further analysis
showed the complex to have a molecular mass of ∼ 100 kDa,
consistent with the expected heterotetrameric structure of GINS,
and to be globular with a ring- or C-shape (discussed below) [4].
To test whether GINS was required for chromosome
replication, the ability of Sld5-depleted Xenopus egg extracts to
replicate exogenously added sperm chromatin was assayed by
monitoring the incorporation of labelled nucleotide analogues
into chromosomal DNA in Sld5-depleted and mock-depleted
S-phase egg extracts [4]. Depending on the exact method used,
either little or no incorporation of the labelled nucleotides was
seen in extracts depleted for Sld5, showing that Sld5, and
most probably the other subunits of GINS, are required for
efficient replication initiation in Xenopus egg extracts. Replication
activity could be restored by adding-back recombinant GINS.
Further analysis using the egg extract system showed that GINS
became associated with sperm chromatin sometime after pre-RC
formation in an S-CDK-dependent manner and dissociated from
chromatin as replication progressed [4]. The timing of GINS
association with chromatin was indistinguishable from that of
Cdc45 and, indeed, the two binding events are interdependent:
GINS does not become chromatin-associated in Cdc45-depleted
extracts and, as in budding yeast, Cdc45 does not become
chromatin-associated in Sld5-depleted extracts. In both cases,
lack of chromatin binding leads to a failure to load the replicative
polymerases Polα and Polε [4]. Cut5 was also required for GINS
and Cdc45 chromatin loading (as in fission yeast, Cut5 is the
orthologue of Dpb11; see Figure 1).
Human
Genes encoding four subunits of human GINS were identified
by comparative sequence analysis and their products shown
to form a ∼ 100 kDa heterotetrameric complex with 1:1:1:1
stoichiometry when expressed in recombinant form [43,44]. In
addition, monoclonal antibodies raised against Psf2 have been
used to pulldown the GINS complex from human cell extracts
[45]. As in yeast, the complex is present throughout the cell
cycle, but varies in level in different cell types with significantly
higher concentrations seen in immortalized and transformed cells,
probably due, at least in part, to simultaneous transcriptional
up-regulation of the four genes [45]. Even in untransformed
cell lines, such as Wi38 human fibroblasts, there are ∼ 100 000
GINS complexes per cell. As cells leave the cell cycle due
to contact inhibition of growth, these levels drop markedly, so
that the four GINS subunits are undetectable in cell extracts
prepared from G0 non-cycling cells. Protein turnover probably
has an important influence here, as the half-lives of the GINS
protein are in the 6–8 h range [45]. While the level of GINS is
constant throughout the cell cycle, a significant fraction of the
complex becomes chromatin-associated during S-phase, as is
the case in yeast and Xenopus. This chromatin-associated material
co-purifies with MCM and Cdc45 from extracts prepared from
nuclease-treated S-phase chromatin [45], consistent with the
existence of CMG complexes in human cells. siRNA (small
Structure and function of the GINS complex
interfering RNA) is a powerful tool for dissecting gene function
in mammalian cells [46,47]. Transfection of siRNAs directed
against the GINS subunits individually into HeLa cells reduces
GINS levels to <10 % of that seen in control transfected cells and
inhibiting the expression of any one subunit of the complex results
in degradation of the others [45]. Cell proliferation and DNA
replication (measured by uptake of labelled precursors and flow
cytometry) is significantly impaired in HeLa and HDF (human
dermal fibroblast) cells transfected with GINS-targeted siRNAs,
with defects seen at both the initiation and elongation stages
of replication and the apparent accumulation of DNA damage
[45,48]. Pre-RC assembly is unaffected in HDF cells, as judged by
the loading of MCM on to chromatin, whereas Cdc45 recruitment
to chromatin is delayed and the maximal level of chromatin-bound
Cdc45 reduced [48]. Further studies suggest that GINS depletion
in HeLa human cells leads to mitotic defects also [14], although
this is not seen in untransformed HDF cells [48]; this will be
discussed further shortly.
Recombinant human GINS has been used in both biochemical
and structural studies [8–10,43,44]. In EMSAs (electrophoretic
mobility-shift assays), GINS has been shown to be able to
bind to DNA with a preference for single-stranded molecules
or molecules containing stretches of ssDNA (single-stranded
DNA), including bubble structures [43]. To date, this remains the
only evidence for direct GINS–DNA interactions. Using surface
plasmon resonance and co-immunoprecipitation, human GINS
has been shown to interact directly with the dimeric primase
component of Polα-primase, the enzyme responsible for synthesis
of the RNA–DNA primers that initiates synthesis of the leading
strand and each Okazaki fragment on the lagging strand [44].
A similar interaction has been observed with the archaeal GINS
and primase proteins (see below) [12]. In addition to binding
to primase, human GINS also appears capable of stimulating its
activity: a 10-fold stimulation is seen when GINS is present in a
200-fold molar excess over primase in assays using singly primed
single-stranded M13 DNA [44].
GINS is a component of the CMG complex
The experiments described above clearly place GINS at the
replication fork, performing an essential role for the initiation and
elongation stages of replication, but what does the GINS complex
actually do? To date, no stand-alone enzymatic activity has been
ascribed to GINS. Instead, GINS appears to perform its function
as part of the three-component CMG complex. CMG, an acronym
for Cdc45–MCM–GINS, was first identified by Moyer et al. [29]
in the course of purifying high molecular mass complexes from
Drosophila embyro extracts that contained the essential Cdc45
protein [29]. In these experiments, Cdc45 co-purified with ten
other proteins in a complex with an apparent molecular mass
of 700 kDa – the six subunits of the MCM helicase (Mcm2–
Mcm7) and the four subunits of Drosophila GINS (Sld5, Psf1,
Psf2 and Psf3). Only ∼ 5 % of the total amount of Cdc45 and
GINS in the extract was in the CMG complex and even less of
the MCM helicase (∼ 1 %), suggesting that assembly of the CMG
might be regulated by, for example, post-translational regulation
of one or other of the components. Nevertheless, as isolated
from Drosophila embryo extracts, the CMG complex is stable
and, more importantly, biochemically active as a DNA helicase,
which raised the possibility that the CMG complex was the elusive
replicative helicase in eukaryotic cells [29].
Further evidence for CMG being the replicative helicase comes
from studies of plasmid replication in Xenopus egg extracts. In
these assays, plasmid DNA is incubated in an extract (termed
493
HSS for high-speed supernatant) derived from egg cytoplasm
to promote pre-RC formation, after which a concentrated NPE
(nucleoplasmic extract) is added to bring about replication
initiation [49]. Initiation occurs independently of DNA sequence
in this system, restricting detailed analysis of protein–DNA
interactions at replication origins. To circumvent this restriction,
Pacek et al. [50] synthesized plasmid DNA that contained a
biotin group at a single site, to which a streptavidin moiety
was bound. The presence of the biotin–streptavidin pairing is
sufficient to bring about replication fork pausing at the labelled
site, allowing ChIP to be used to monitor the presence or absence
of individual target proteins in the paused replisome. In addition,
it had previously been shown that adding the DNA polymerase
inhibitor aphidicolin results in the uncoupling of DNA synthesis
from DNA unwinding in these reactions [49]; the uncoupled
DNA helicase continues to unwind the duplex while nascent DNA
synthesis ceases. Under these conditions, the helicase alone would
be expected to be present at the pause site on the plasmid. Using
ChIP, it was shown that the MCM helicase was indeed present at
the pause site, as were Cdc45 and the GINS complex. All three
replicative DNA polymerases (α, δ and ε) were also present but
not in the presence of aphidicolin, when synthesis and unwinding
are uncoupled. It is highly likely that the MCM, Cdc45 and GINS
proteins seen at the pause site are in the form of the CMG.
GINS is a component of the replisome progression complex
In addition to making up the CMG complex, it is now clear that
Cdc45, MCM and GINS are part of a larger complex present in
S-phase at the replication fork. Tandem affinity purification of
GINS from budding yeast cell extracts brings down, in addition
to Cdc45 and MCM, a specific set of proteins that mediate or
regulate replication fork progression [6,30]. This large set of proteins has been designated the RPC. The list of co-purifying
proteins identified in these experiments is long, but all six subunits
of MCM are present, as is Cdc45. Also present are Tof1 and Csm3
(which allow forks to pause when encountering protein barriers
to replication), Mrc1 (which mediates checkpoint signalling in
response to fork stalling), Ctf4 (which interacts with Polαprimase and is required for the establishment of sister chromatid
cohesion), Top1 (topoisomerase I, which is thought to remove
positive supercoils generated by DNA unwinding ahead of the
fork), Mcm10 (which also interacts with Pol α-primase), FACT
(the histone chaperones Spt16 and Pob3) and, under certain
purification conditions, Pol α-primase itself [7]. Interestingly,
evidence from budding yeast suggests that each RPC contains
only a single MCM helicase at its core and therefore only a
single CMG complex, and that GINS plays an essential role in
maintaining the interaction of MCM with Cdc45 within the CMG
[6]. While this does not rule out the possibility that the RPC
may contain multiple GINS complexes, performing additional
functions other than stabilizing the MCM–Cdc45 interaction in
the CMG, evidence for this is currently lacking.
GINS STRUCTURES
Primary structures
As noted above, the four subunits of eukaryotic GINS are distantly
related to one another at the protein sequence level and most
probably descended from a common ancestor, that is, they are
paralogous. This was first shown by Koonin and colleagues [11]
who identified GINS homologues in archaea also (discussed in
detail below). Interestingly, the four eukaryotic GINS proteins can
c The Authors Journal compilation c 2010 Biochemical Society
494
S. A. MacNeill
be recognized as being of two types that differ in the arrangement
of conserved protein sequences (Figure 2A). Two regions of
protein sequence conservation can be identified in Sld5 and Psf1
that correspond, as we shall see shortly, to distinct domains
in the three-dimensional structures of the proteins. In Psf2 and
Psf3, both of these domains (termed the A- and B-domains) are
present, but their order in the primary sequence of the proteins
is reversed relative to that seen in Sld5 and Psf1: the larger
N-terminal A-domain in Sld5 and Psf1 is C-terminally located in
Psf2 and Psf3, whereas the smaller C-terminal B-domain of Sld5
and Psf1 is N-terminally located in Psf2 and Psf3 (Figure 2A).
Presumably the four genes encoding the eukaryotic GINS subunits
evolved from an ancestral gene by two rounds of gene duplication
with the first round being followed by a rearrangement (circular
permutation) of one of the two duplicated genes that reversed the
domain order [11]. In archaea, proteins with A-B domain order
predominate (see Figure 2B), perhaps suggesting that the ancestral
GINS protein was of this type, although other interpretations are
possible, particularly if there have been archaeal lineage-specific
losses of genes encoding B-A-type proteins.
Crystal structures
In independent studies, three groups have solved crystal structures
of human GINS with resolutions varying from 2.3 Å to 3.2 Å
(1 Å = 0.1 nm) [8–10]. The structures (PDB codes 2E9X, 2EHO
and 2Q9Q) are highly similar to one another, but none shows
the entire structure of the complex. The overall shape of the
heterotetrameric complex has been variously described as
resembling a trapezium [9], an elongated spindle with a visible
central hole [8] or of being of elliptical shape [10]. Figure 3
shows the overall structure of the complex (an interactive
version of this can be seen at http://www.BiochemJ.org/bj/425/
0489/bj4250489add.htm), while for clarity, Figure 4 shows views
of individual subunits and domains.
The overall dimensions of the complex are ∼ 110 Å × 60
Å × 60 Å, the central hole is perhaps 5–10 Å in diameter [8–10].
The four subunits are arranged in two layers: Sld5 and Psf1 form
the top layer and Psf2 and Psf3 the bottom layer. This arrangement
of subunits is consistent with observations made with recombinant
Xenopus and human GINS proteins [4,43], where Sld5, Psf1 and
Psf2 have been shown to be able to form a stable trimer in the
absence of Sld3, and Sld5 and Psf2 a stable dimer in the absence
of Psf1 and Psf3, although it should be noted that there is no
evidence that such sub-complexes exist in vivo.
All four subunits are related to one another in structure, as
they are in sequence, and the arrangement of subunits creates
a vertical pseudo 2-fold axis in the centre of the complex
(indicated by the broken line in Figure 3). Consistent with the
bioinformatic analysis discussed above, each subunit is composed
of two distinct domains joined by a linker region of variable
length: the predominantly α-helical A-domain and the largely
β-stranded B-domain. As indicated above, the A-domain is
N-terminal in Sld5 and Psf1 but C-terminal in Psf2 and Psf3.
The A-domains comprise five α-helices in Sld5, Psf1 and Psf3
and four α-helices in Psf2 (a β-strand replaces one of the five
helices seen in the other four subunits) and form an arch shape
(this is particularly noticeable in Sld5 and Psf1; see Figure 4A)
[8–10]. Heterodimerization of the A-domains of the top layer
subunits Sld5 and Psf1, and of the bottom layer subunits Psf2 and
Psf3, creates the vertical interface in the tetrameric complex and
creates a 2-fold pseudo-symmetrical axis (indicated in Figure 3).
Key to the structure of each individual A-domain is an intrasubunit
bidentate hydrogen bond formed, in the otherwise markedly
c The Authors Journal compilation c 2010 Biochemical Society
Figure 2
Domain structures and genomic context
(A) Schematic representation of the domain structures of eukaryotic and archaeal GINS
proteins. Note that the length of the linker region (shown as a thin line) between the
A- and B-domains is highly variable. In certain haloarchaeal species, for example, this
extends to 150–200 amino acids [66]. (B) Chromosome context of GINS genes from
14 archaeal species representative of the major orders of the Crenarchaeota (Sso, S.
solfataricus ; Ape, Aeropyrum pernix ; Pae, Pyrobaculum aerophilum ), the Euryarchaeota
(Mth, Methanothermobacter thermoautotrophicus ; Mma, Methanococcus maripaludis ; Tac,
Thermoplasma acidophilum ; Afu, Archaeoglobus fulgidus ; Mac, Methanosarcina acetivorans ;
Mhu, Methanospirillum hungatei ; Hma, Haloarcula marismortui ; Pfu, Pyrococcus furiosus ; Neq,
Nanoarchaeaum equitans ), the Thaumarchaeota (Thaum.) (Csy, Cenarchaeum symbiosum ) and
the Korarcheaota (Kar.) (Kcr, Korarchaeum cryptofilum ). Entrez GeneIDs can be found in the
Supplementary Table S1 (at http://www.BiochemJ.org/bj/425/bj4250489add.htm). See [54,55]
for further information on additional linked genes encoding proteins involved in translation.
Note that the ORF (open reading frame) lengths are not shown to scale and that overlapping
ORFs are not shown as such. See text for further details.
Structure and function of the GINS complex
Figure 3
495
Structure of the human GINS complex
Schematic representation of the three-dimensional structure of the human GINS complex showing the overall shape of the complex, the relative positioning of the four subunits (Sld5, yellow; Psf1,
green; Psf2, cyan; Psf3, purple) and the pseudo 2-fold axis of symmetry (vertical line). Sld5 and Psf1 form the top layer of the complex, Psf2 and Psf3 the bottom layer. The structure on the right-hand
side is rotated 90◦ relative to that on the left-hand side and shows the position of the B-domain of Sld5. The structure shown was derived from PDB entry 2E9X [9] and drawn using MacPyMOL
(DeLano Scentific). An interactive three-dimensional version of the structure can be found at http://www.BiochemJ.org/bj/425/0489/bj4250489add.htm
hydrophobic environment, between a highly conserved arginine
residue in the α3 helix and a highly conserved glutamate residue
in the α5 helix. Replacement of the arginine residue in fission
yeast Psf2 with a lysine residue is the cause of the conditionallethal (temperature-sensitive) phenotype of the psf2-209 allele
[36], highlighting the importance of the hydrogen bond for the
Psf2 structure and function.
The smaller B-domain comprises an extended N-terminal part
which includes a single α-helix followed by a globular core which
comprises up to five β-strands, arranged in two antiparallel
β-sheets, and a single α-helix (shown for Sld5 and Psf2 in
Figure 4B) [8–10]. Although the A-domain fold is reported to
be distantly related to part of the structure of the cytoskeletal
protein spectrin [9], no domains similar to the B-domain have
been found.
In the tetramer, the B-domains of Psf2 and Psf3 create, through
their interactions with the A-domain α1 and α3 helices of Sld5
and Psf1, the horizontal interface of the top and bottom layers
(Figure 3) [8–10]. The effects of disrupting this interface can
be seen with the budding yeast psf1-1 allele [5]; this mutation
maps to an arginine residue in the Psf1 α3 helix that would
normally be expected to protrude into a hydrophobic pocket
in the B-domain of Psf3. Replacement of the arginine residue
with glycine renders the mutant protein temperature-sensitive.
In contrast with the N-terminal B-domains of Psf2 and Psf3, the
C-terminal B-domain of Sld5 is located between the A-domains of
Sld5 and Psf2 on the front surface of the complex (see Figure 3,
right-hand panel). Deletion of the Sld5 B-domain weakens the
horizontal interface between the top and bottom layers of the
complex without affecting the vertical interfaces between Sld5
and Psf1 or Psf2 and Psf3 [9].
Absent from all three crystal structures is the C-terminal
B-domain of Psf1. In two of three cases [9,10], the B-domain was
removed from the complex prior to crystallization, whereas in the
third [8], the B-domain was present in the crystallized material
but was invisible in the electron density maps. Interestingly,
biochemical experiments indicate that this domain does not
interact stably with a tetrameric complex comprising C-terminally
truncated Psf1 and full-length Sld5, Psf2 and Psf3, prompting
the suggestion that it is tethered to the core complex by its
linker region only [9]. Although not required for core complex
formation, the Psf1 B-domain does have an essential function,
however, and may be involved in protein–protein interactions
with other components of the replication machinery. Addition of
recombinant human GINS to egg extracts depleted of endogenous
GINS restores full replication activity, but only when Psf1
is present in full-length; human GINS complexes containing
C-terminally truncated Psf1 are incapable of replicating sperm
chromatin, even when recombinant Psf1 B-domain is added
in trans. The truncated complexes fail to stably associate
with chromatin and Cdc45 chromatin binding is also greatly
reduced, consistent with previous observations of interdependent
chromatin binding of GINS and Cdc45 in Xenopus extracts [4].
Polε also fails to bind chromatin, whereas ORC and MCM loading
(pre-RC formation) is unaffected (see Figure 1).
Note that, in addition to the Psf1 B-domain, the last few residues
of Psf2, including putative phosphorylation sites for the ATM
(ataxia-telangiectasia mutated) and ATR (ATM and Rad3 related)
kinases, are also missing from all three structures (discussed
further below).
EM (electron microscopy) structures
Prior to the publication of the crystal structures of the human
GINS complex, two EM studies of GINS structure were reported.
In the original report on the identification and characterization
of Xenopus GINS, transmission EM images of rotary-shadowed
GINS complexes (reproduced in Figure 5) were presented that
appear to show a ring- or C-shaped structure with an estimated diameter of 100 Å and a central channel or hole with an estimated
diameter of 40 Å [4]. A second EM study, this time of the
recombinant human GINS complex, showed a C-shaped complex
of broadly similar dimensions [43]. Neither of these EM structures
is entirely consistent with the crystal structures; in particular, the
size of the central channel is significantly smaller in the crystal
structures and may not be large enough to accommodate singleor double-stranded DNA, as was previously suggested [4,43].
The reasons for these differences are unclear, but may reflect the
limitations of the low-angle rotary shadowing technique used to
prepare the GINS complexes for EM analysis [9]. Should it exist,
c The Authors Journal compilation c 2010 Biochemical Society
496
S. A. MacNeill
Figure 5
EM of Xenopus GINS
Upper panels: transmission electron micrographs of rotary-shadowed recombinant Xenopus
GINS complexes apparently showing ring-like structures with central cavities. Reproduced from
c 2003 Cold Spring Harbor Laboratory Press. Lower panel: crystal
[14] with permission; structure of human GINS orientated to show the central cavity and the location of the Psf3
N-terminal peptide. The image was drawn with MacPyMOL (DeLano Scientific) using PDB file
2E9X [9]. The colour scheme is the same as that used in Figure 3.
ARCHAEAL GINS
Figure 4
The human GINS subunits possess a common fold
(A) Crystal structures of the four individual human GINS subunits. The N-terminal A-domains
of Sld5 and Psf1 form an arch shape with the C-terminal B-domain positioned above (seen for
Sld5 only, as the Psf1 B-domain is not present in the solved structure). The arch shape of Psf2
and Psf3 is less apparent, but is formed by the A-domain helices and resembles that of Sld5 and
Psf1. In Psf2 and Psf3, however, the B-domain is located at one end of the arch, at the N-termini
of the proteins. (B) The isolated globular cores of the B-domains of Sld5 (left-hand side) and
Psf2 (right-hand side). The structures shown are derived from PDB 2E9X [9] and drawn using
MacPyMOL (DeLano Scientific).
the function of the central channel is unknown; a possible
role for a short peptide from the N-terminus of Psf3 in gating
access to the channel has been proposed [8], but what might be
being gated is unclear.
c The Authors Journal compilation c 2010 Biochemical Society
The archaea comprise the third domain of life on Earth
and possess chromosomal replication machinery that resembles
a highly simplified form of that found in eukaryotic
cells [51]. This apparent simplicity, coupled with the ease
with which archaeal replication factors, particularly those
derived from hyperthermophilic archaeal organisms, can be
expressed and purified in recombinant form, has led to
the adoption of the archaea as an important model system
for eukaryotic chromosome replication. At least three major
archaeal phyla have been identified: the Crenarchaeota (which
include only hyperthermophilic organisms), the Euryarchaeota
(includes methanogens, halophiles, thermoacidophiles and
some hyperthermophiles) and the Thaumarchaeota [52]. The
Korarchaeota may represent a fourth phylum [53]. Among
the replication proteins relevant to the present review that are
conserved between the archaea and eukaryotes are ORC and
Cdc6 (in archaea, the functions of these two factors appear to
reside in one protein which has homology with both ORC and
Cdc6), the MCM helicase (homohexameric in archaea, rather
than heterohexameric), primase and GINS [51].
Identification and distribution of archaeal GINS
homologues
Archaeal GINS homologues were first identified by bioinformatic
methods [11] and are found encoded by all archaeal species
Structure and function of the GINS complex
whose genomes have been sequenced to date (S. A. MacNeill,
unpublished work). As in eukaryotic cells, two types of GINS
proteins are recognizable on the basis of the organization of the
conserved A- and B-domains (shown in Figure 2A). All archaeal
organisms encode a protein with an A-B domain order resembling
eukaryotic Sld5 and Psf1. These have been named Gins15 [12]
or Gins51 [13]; in the present review, the latter nomenclature is
used, consistent with the 5-1-2-3 origin of the term GINS (goichi-ni-san). In many cases, the gene encoding the Gins51 protein
is found adjacent in the genome to the gene encoding PriS, the
small subunit of primase, the enzyme responsible for synthesis
of the RNA primer at the 5 end of each Okazaki fragment on
the lagging strand as well as at the 5 end of the leading strand
at the replication origin, or the gene encoding the polymerase
processivity factor and sliding clamp PCNA (proliferating-cell
nuclear antigen), or both. Figure 2(B) shows the genome context
of genes encoding Gins51 proteins in species representative of the
individual orders of the four putative archaeal phyla. In seven of
14 species shown, the gene encoding the primase subunit PriS is
located immediately upstream of that encoding Gins51, whereas
in three species, the gene encoding PCNA is found immediately
upstream. In the three crenarchaeal organisms, all three genes
are linked, whereas in the sole sequenced korarchaeal species,
Korarchaeum cryptofilum, the gene encoding the large subunit of
primase is also present (Figure 2B). In the remaining four species,
the genes are not linked. Further analysis of gene context has
established an intriguing link between replication and translation,
discussion of which lies outside the scope of the present review
[54,55].
Unlike the Gins51 proteins, Gins23 proteins, with their
characteristic B-A domain order similar to Psf2 and Psf3, have
been found encoded by some, but not all, archaeal species.
Most of the euryarchaea, including the methanogenic archaea
(represented by the species indicated by Mth, Mma, Mac and
Mhu in Figure 2B), the haloarchaea (Hma) and the nanoarchaea
(Neq), appear to lack Gins23. This may simply indicate that
the sequences of the Gins23 proteins in these organisms have
diverged to the point where they can no longer be detected on
the basis of amino-acid-sequence similarity: the known archaeal
Gins23 proteins are already highly divergent. However, if Gins23
proteins are truly absent, intriguing questions are raised regarding
the structure and stability of the GINS complex. As we shall see
shortly, where both Gins51 and Gins23 proteins are present, the
archaeal GINS complex is a tetramer made up of homodimers
of each protein [12,13]. Given the important role for the
N-terminal B-domains of Psf2 and Psf3 in forming the horizontal
interface of the eukaryotic GINS complex (shown in Figure 3),
it may be the case that in the absence of Gins23, the Gins51
protein is physically incapable of assembling into a stable twolayered, tetrameric structure like that found in eukaryotes and
presumably also in archaea that encode both Gins51 and Gins23.
Thus the GINS complex in these organisms may be dimeric. Since
Gins51 proteins can be capable of homodimer formation when
expressed in recombinant form (S. A. MacNeill, unpublished
work), this question will only be answered by purification of GINS
complexes from native protein extracts prepared from archaeal
organisms that apparently lack Gins23. Interestingly, two of the
species that possess genes encoding both Gins51 and Gins23
(Cenarchaeum symbiosum and K. cryptofilum, indicated as Csy
and Kcr in Figure 2B) have been proposed to be members of deepbranching, ancient archaeal orders [52,53]. If this is true, it implies
that the gene encoding the Gins23 protein was present in the last
common ancestor of the extant species and then lost at a later
stage of evolution, after the archaeal and eukaryotic lineages had
diverged.
497
Biochemical activity
To date, two GINS complexes from evolutionarily distinct
hyperthermophilic species, the crenarchaeote Sulfolobus
solfataricus and the euryarchaeote Pyrococcus furiosus, have
been characterized biochemically [12,13]. Both species encode
both Gins51 and Gin23 proteins (Figure 2B). In S. solfataricus,
Gins23 has been shown to interact with the non-catalytic
N-terminal domain of the MCM helicase in yeast two-hybrid
assays (indeed, this was how the gene was isolated, in a twohybrid screen with MCM as the bait) and by co-immunoprecipitation from S. solfataricus cell extracts [12]. The genes encoding
these two proteins overlap on the chromosome by 11 bp
(Figure 2B). Gins23 also interacts in yeast two-hybrid assays and
in vitro pull-down experiments with both subunits of primase.
Purification of Gins23 from native protein extracts identified two
further interacting factors, one of which was shown to be a GINS
family protein and named Gins51. The gene encoding Gins51
is immediately adjacent in the chromosome to that encoding the
small catalytic subunit of primase (Figure 2B).
When co-expressed in recombinant form, S. solfataricus
Gins23 and Gins51 form a tetrameric GINS complex comprising a
dimer of dimers. This complex presumably adopts a structure very
similar to that seen with the human GINS complex (Figure 3), but
with the difference that the top and bottom layers are composed
of Gins51 and Gins23 homodimers respectively, rather than
Sld5–Psf1 and Psf2–Psf3 heterodimers. All other aspects of the
structure of the complex, such as the role of the Psf2 and Psf3
B-domains in creating, through their interactions the A-domains
of Sld5 and Psf1 respectively, the horizontal interface between the
top and bottom layers, might be expected to be preserved in the
S. solfataricus GINS structure.
The other protein that co-purified with GINS displayed
sequence similarity to the DNA-binding domain of the bacterial
RecJ family of ssDNA exonucleases [12]. The RecJdbh protein
(RecJ DNA-binding domain homologue) remained associated
with the GINS protein through eight chromatographic steps,
suggesting a very stable interaction, but its function is unknown
[12]. Nevertheless, it is tempting to speculate that the potential
ability of the RecJdbh to bind to ssDNA may be significant in the
context of the likely role and location of the GINS complex at
the replication fork.
The other archaeal GINS to be analysed biochemically is that
of P. furiosus [13]. When expressed in recombinant form, the
P. furiosus Gins51 and Gins23 proteins also form a tetrameric
GINS complex in a 2:2 molar ratio. As seen with the S. solfataricus
proteins, the P. furiosus Gins23 protein interacts in a two-hybrid
assay with the MCM helicase, and the two proteins can be
co-immunoprecipitated from cell extracts. The corresponding
genes are located next to one another on the chromosome also
(Figure 2B). Two-hybrid analysis also shows the Gins51 protein
to interact with the P. furiosus Orc1/Cdc6 homologue, and
ChIP studies using anti-Gins23 sera show the GINS complex to
associate with the P. furiosus replication origin in exponentially
growing, but not stationary phase, cells [13].
Interestingly, recombinant P. furiosus GINS has been shown
in in vitro assays to be able to stimulate MCM ATPase and
helicase activities when present at up to a 4-fold molar excess
over MCM [13], echoing the situation in eukaryotic cells with
the CMG complex [29]. Unlike eukaryotic cells, however, there
are no indications as yet that the archaea have the potential
to encode a homologue of Cdc45, the third component of the
CMG complex. It is possible that this protein is present in
archaeal cells but that it has diverged in sequence to the point
where it cannot be detected by even the most thorough database
c The Authors Journal compilation c 2010 Biochemical Society
498
S. A. MacNeill
searching. Alternatively, Cdc45 may only have evolved after the
eukaryotic and archaeal lineages diverged in evolution. In this
scenario, archaeal organisms might make do without a Cdc45like function or that function might be provided by an altogether
different protein, one such as RecJdbh perhaps. Taken together,
the results from S. solfataricus and P. furiosus [12,13] place the
archaeal GINS complex at the heart of the replication apparatus,
making contact with Orc1/Cdc6 at origins and with primase and
MCM in the moving replisome, and in doing so underline the
credentials of the archaea as simplified models for understanding
the complexities of chromosome replication in eukaryotes.
LINKS TO MITOSIS
In eukaryotic cells, evidence exists pointing to a possible role
for GINS in chromosome segregation that appears distinct from
its role in S-phase [14]. The human Survivin protein forms a
complex (called the CPC or chromosomal passenger complex)
with the mitotic protein kinase Aurora B, the inner centromere
protein INCENP and a protein called borealin, that apparently
acts to co-ordinate multiple events during mitosis and cytokinesis
[56]. The CPC localizes on the spindle in mitotic metaphase
before relocating to the spindle midzone in anaphase. Survivin is
phosphorylated by Aurora B and may have a role in regulating the
activity of the kinase, which has itself been implicated in a number
of processes including monitoring spindle tension, ensuring
correct chromosome bi-orientation and spindle disassembly
[56].
In fission yeast, the psf2+ gene was identified as a multicopy
suppressor of a temperature-sensitive mutation, bir1-46, in the
Survivin homologue Bir1 [14]. Overexpression of the Psf2 protein
also rescues additional bir1-46 mutant phenotypes including
TBZ sensitivity (TBZ is a microtubule poison), suggesting a
possible role for the protein in mitosis or cytokinesis. In support
of such a role, a significant proportion of psf2 cells display
chromosome missegregation phenotypes in mitosis, such as
unequal nuclear division or lagging chromosomes. The same is
true of cells in which psf2+ mRNA is depleted by promoter shutoff, in cells carrying the temperature-sensitive psf2-209 allele
described above, and in sld5 cells. In addition, the localization
of Bir1 in anaphase and telophase is disrupted in sld5 and
psf2 cells. Human Psf2–EGFP [EGFP is enhanced GFP (green
fluorescent protein)] fusion proteins localize in part to the spindle
in metaphase and anaphase and to the midzone in cytokinesis,
and, in transformed human HeLa cells, siRNA depletion of Psf2
also causes defects in the maintenance of nuclear morphology and
chromosome missegregation. Fission yeast Psf2–GFP may also
localize to the mitotic spindle [14].
Taken together, these results suggest that GINS may have an
important role in mitosis and cytokinesis in eukaryotic cells.
The key question, however, is to what extent is such a role
distinct and separable from the S-phase function of GINS? Are
the observed mitotic defects simply a consequence of trying to
separate incompletely replicated chromosomes? Attempts have
been made in fission yeast to address this issue, by waiting until
S-phase is complete before depleting or inactivating GINS in G2
and then following the effects on the subsequent mitosis, but
without clear-cut results [14]. In addition, a more recent study,
this time using untransformed HDFs rather than HeLa cells,
failed to observe aberrant mitotic phenotypes upon simultaneous
siRNA depletion of Psf1 and Psf2 [48]. The explanation
for this difference in behaviour is unclear, but begs further
investigation.
c The Authors Journal compilation c 2010 Biochemical Society
GINS FUNCTION IN VERTEBRATE DEVELOPMENT
Several studies have appeared that characterize the requirement
for GINS during vertebrate development [15–19]. In Xenopus,
expression patterns of SLD5, PSF1, PSF2 and PSF3 have been
analysed by in situ hybridization during embryonic development
[18]. These studies showed the expression patterns of the four
genes to be generally coincident, suggesting that they share some
common regulation. Interestingly, high levels of expression do
not always correspond with high rates of cell proliferation, again
possibly hinting at roles for GINS outside of S-phase. Injection of
anti-sense oligonucleotides (morpholinos) into Xenopus embryos
to knockdown PSF2 function results in impairment of a number
of developmental processes, among which eye development has
been studied in detail [19].
In adult mice, PSF1 expression is readily detected in
activity-proliferating tissues, such as bone marrow and thymus,
and also in the reproductive tissues, the testis and ovary.
Immunohistochemical studies of testis show the Psf1 protein
to be located in immature cell populations, including the
blastocysts and haematopoietic progenitor cells [16]. Targeted
disruption of one Psf1 allele in mice results in the expected
reduction of Psf1 transcript levels, but has no gross phenotypic
consequences. However, crossing heterozygous Psf1+/− mice
does not produce viable Psf1−/− offspring, with Psf1−/− embryos
failing to develop post-implantation beyond embryonic stage
E5.5 (where E is embryonic day), when blastocysts are present.
It is likely that viability through the early stages of embryo
development, up to E5.5, is assured by the presence of
maternal Psf1 transcripts. As the embryos reach stage E5.5,
Psf1 expression from the maternal mRNA is insufficient for
continued cell growth [16]. In culture, Psf1−/− blastocysts display
defects in cell proliferation and impaired DNA replication, as
judged by failure to incorporate the nucleotide analogue BrdU
(bromodeoxyuridine) into DNA. Similar phenotypes are seen in
homozygous Cdc45−/− embryos [57]. Although the heterozygous
Psf1+/− embryos are viable and do not display gross phenotypic
defects, closer investigation of the processes of haemopoiesis
has revealed that Psf1 haploinsufficiency leads to impaired
stem-cell proliferation in response to haematopoietic stress and
consequent reduced survival [17]. Underlining the importance
of GINS in normal development, it has also been reported that
PSF2 expression is significantly elevated in certain types of
human hepatic carcinomas [15]. Whether this contributes directly
to tumorigenesis is unclear at present, but as mutations in
MCM4 and down-regulation of MCM2 have already been shown
to cause cancer in mice [58–60], it would not be surprising
if deregulation or mutation of GINS were to have similar
consequences.
REGULATION OF GINS ACTIVITY
The activities of various components of the replication machinery
are tightly regulated, to ensure once-per-cell-cycle replication, for
example, or to allow the replication machinery to respond to the
ever-present threat of DNA damage, and recent results indicate
that this is also likely to be the case for GINS [20]. Human Psf2
was identified in an extensive proteomic-based survey of proteins
phosphorylated by the ATM and ATR protein kinases [20], two
key mediators of the DNA-damage response with overlapping, but
non-redundant, functions [61,62]. In the case of Psf2, two sites of
phosphorylation were identified: Thr180 and Ser182 . ATM and ATR
recognize serine and threonine residues in SCDs (SQ/TQ cluster
domains) [63]: Thr180 and Ser182 are found in the sequence SQTQ,
Structure and function of the GINS complex
close to the C-terminus of Psf2 (and absent from all three crystal
structures of the human GINS complex) [8–10].
At present it is not known what effect these phosphorylation
events have on GINS function, nor is it known how they contribute
to the cellular DNA-damage response. Neither phosphorylation
site is particularly well-conserved across evolution. Budding yeast
Psf2, for example, does not contain any SQ/TQ motifs, although
fission yeast Psf2 does have an SQ close to its C-terminus. It
would be interesting, and relatively straightforward technically,
to determine whether mutating this site affected the ability of the
cells to deal with the DNA damage in S-phase.
Another aspect of GINS function in higher eukaryotes that
has been little investigated is the potential role for differential
expression or alternative splicing in generating different isoforms
of the GINS proteins. In a recent study, differences in PSF1
transcript structure in testis were described that could potentially
give rise to the expression of N-terminally truncated Psf1 proteins
[64]. Direct biochemical evidence for the existence of such
proteins is lacking, however, so it remains to be seen whether
variants of this type have any part to play in GINS function.
CONCLUSIONS AND PERSPECTIVES
Since its discovery in 2003 [3–5], GINS has moved rapidly to
centre stage in our attempts to fully understanding the workings
of the chromosome replication machinery in eukaryotic and
archaeal cells. GINS is undoubtedly a key component of the
likely replicative helicase, the CMG complex [29,50], but what
exactly does GINS contribute to the activity of the CMG? What
is its biochemical function? The structure of the GINS complex
in isolation [8–10] offers no clues to this. The requirement for
GINS for replication elongation suggests an intimate association
with processive DNA unwinding, but more detailed speculation
on GINS function is hampered by our lack of understanding of the
mechanism of the MCM helicase itself. As discussed elsewhere
[65], various models have been proposed for how MCM might act
at the fork, largely depending on whether the protein is present
as a hexamer or a double hexamer, and whether the ring-shaped
MCM complex encircles single- or double-stranded DNA. The
elegant demonstration that the RPC contains only a single MCM
complex [6] probably resolves the first of these issues, but the
single-stranded versus double-stranded DNA debate continues.
One intriguing model posits that the MCM hexamer is not actually
a helicase at all, i.e. that it is not intrinsically able to unwind duplex
DNA, but simply a double-stranded DNA translocase [65]. In this
model, duplex DNA is pumped in an ATP-dependent manner
through the central channel of the MCM hexamer without being
unwound. However, exit from the channel for duplex DNA is
blocked by the presence of an unknown protein and it is some part
of the structure of this additional factor (termed the ploughshare
after the main cutting blade of a plough) that sterically separates
the two strands. Could GINS harbour the ploughshare? The
ability of recombinant human GINS to bind preferentially ssDNA
[43] would certainly be consistent with GINS being located
in proximity to unwound DNA emerging from the CMG, but
aside from this, evidence for a precise role for GINS is lacking.
A complete crystal structure of the CMG, ideally complexed
with a synthetic DNA substrate, would be hugely informative
in this regard, but the technical challenges involved in generating
sufficient material in a form appropriate for crystallization are
substantial, and there is no guarantee that crystallization itself
would be successful. Until these challenges are met, further
genetic and biochemical analysis of GINS function in a variety
of experimental systems is required and will no doubt provide
499
important insights into the function of this essential protein
complex.
FUNDING
Work in the author’s laboratory is funded by the Scottish Universities Life Sciences Alliance
(SULSA).
REFERENCES
1 Bell, S. P. and Dutta, A. (2002) DNA replication in eukaryotic cells. Annu. Rev. Biochem.
71, 333–374
2 Dutta, A. and Bell, S. P. (1997) Initiation of DNA replication in eukaryotic cells. Annu. Rev.
Cell Dev. Biol. 13, 293–332
3 Kanemaki, M., Sanchez-Diaz, A., Gambus, A. and Labib, K. (2003) Functional proteomic
identification of DNA replication proteins by induced proteolysis in vivo . Nature 423,
720–724
4 Kubota, Y., Takase, Y., Komori, Y., Hashimoto, Y., Arata, T., Kamimura, Y., Araki, H. and
Takisawa, H. (2003) A novel ring-like complex of Xenopus proteins essential for the
initiation of DNA replication. Genes Dev. 17, 1141–1152
5 Takayama, Y., Kamimura, Y., Okawa, M., Muramatsu, S., Sugino, A. and Araki, H. (2003)
GINS, a novel multiprotein complex required for chromosomal DNA replication in
budding yeast. Genes Dev. 17, 1153–1165
6 Gambus, A., Jones, R. C., Sanchez-Diaz, A., Kanemaki, M., van Deursen, F., Edmondson,
R. D. and Labib, K. (2006) GINS maintains association of Cdc45 with MCM in replisome
progression complexes at eukaryotic DNA replication forks. Nat. Cell Biol. 8, 358–366
7 Gambus, A., van Deursen, F., Polychronopoulos, D., Foltman, M., Jones, R. C.,
Edmondson, R. D., Calzada, A. and Labib, K. (2009) A key role for Ctf4 in coupling the
MCM2–7 helicase to DNA polymerase α within the eukaryotic replisome. EMBO J. 28,
2992–3004
8 Chang, Y. P., Wang, G., Bermudez, V., Hurwitz, J. and Chen, X. S. (2007) Crystal structure
of the GINS complex and functional insights into its role in DNA replication. Proc. Natl.
Acad. Sci. U.S.A. 104, 12685–12690
9 Kamada, K., Kubota, Y., Arata, T., Shindo, Y. and Hanaoka, F. (2007) Structure of the
human GINS complex and its assembly and functional interface in replication initiation.
Nat. Struct. Mol. Biol. 14, 388–396
10 Choi, J. M., Lim, H. S., Kim, J. J., Song, O. K. and Cho, Y. (2007) Crystal structure of the
human GINS complex. Genes Dev. 21, 1316–1321
11 Makarova, K. S., Wolf, Y. I., Mekhedov, S. L., Mirkin, B. G. and Koonin, E. V. (2005)
Ancestral paralogs and pseudoparalogs and their role in the emergence of the eukaryotic
cell. Nucleic Acids Res. 33, 4626–4638
12 Marinsek, N., Barry, E. R., Makarova, K. S., Dionne, I., Koonin, E. V. and Bell, S. D. (2006)
GINS, a central nexus in the archaeal DNA replication fork. EMBO Rep. 7, 539–545
13 Yoshimochi, T., Fujikane, R., Kawanami, M., Matsunaga, F. and Ishino, Y. (2008) The
GINS complex from Pyrococcus furiosus stimulates the MCM helicase activity. J. Biol.
Chem. 283, 1601–1609
14 Huang, H. K., Bailis, J. M., Leverson, J. D., Gomez, E. B., Forsburg, S. L. and Hunter, T.
(2005) Suppressors of Bir1p (Survivin) identify roles for the chromosomal passenger
protein Pic1p (INCENP) and the replication initiation factor Psf2p in chromosome
segregation. Mol. Cell. Biol. 25, 9000–9015
15 Obama, K., Ura, K., Satoh, S., Nakamura, Y. and Furukawa, Y. (2005) Up-regulation of
PSF2 , a member of the GINS multiprotein complex, in intrahepatic cholangiocarcinoma.
Oncol. Rep. 14, 701–706
16 Ueno, M., Itoh, M., Kong, L., Sugihara, K., Asano, M. and Takakura, N. (2005) PSF1 is
essential for early embryogenesis in mice. Mol. Cell. Biol. 25, 10528–10532
17 Ueno, M., Itoh, M., Sugihara, K., Asano, M. and Takakura, N. (2009) Both alleles of PSF1
are required for maintenance of pool size of immature hematopoietic cells and acute bone
marrow regeneration. Blood 113, 555–562
18 Walter, B. E. and Henry, J. J. (2004) Embryonic expression of pre-initiation DNA
replication factors in Xenopus laevis . Gene Expr. Patterns 5, 81–89
19 Walter, B. E., Perry, K. J., Fukui, L., Malloch, E. L., Wever, J. and Henry, J. J. (2008) Psf2
plays important roles in normal eye development in Xenopus laevis . Mol. Vis. 14,
906–921
20 Matsuoka, S., Ballif, B. A., Smogorzewska, A., McDonald, III, E. R., Hurov, K. E., Luo, J.,
Bakalarski, C. E., Zhao, Z., Solimini, N., Lerenthal, Y. et al. (2007) ATM and ATR substrate
analysis reveals extensive protein networks responsive to DNA damage. Science 316,
1160–1166
21 Tanaka, S., Umemori, T., Hirai, K., Muramatsu, S., Kamimura, Y. and Araki, H. (2007)
CDK-dependent phosphorylation of Sld2 and Sld3 initiates DNA replication in budding
yeast. Nature 445, 328–332
c The Authors Journal compilation c 2010 Biochemical Society
500
S. A. MacNeill
22 Zegerman, P. and Diffley, J. F. (2007) Phosphorylation of Sld2 and Sld3 by
cyclin-dependent kinases promotes DNA replication in budding yeast. Nature 445,
281–285
23 Glover, J. N., Williams, R. S. and Lee, M. S. (2004) Interactions between BRCT repeats
and phosphoproteins: tangled up in two. Trends Biochem Sci. 29, 579–585
24 Masumoto, H., Muramatsu, S., Kamimura, Y. and Araki, H. (2002) S-Cdk-dependent
phosphorylation of Sld2 essential for chromosomal DNA replication in budding yeast.
Nature 415, 651–655
25 Tanaka, S., Tak, Y. S. and Araki, H. (2007) The role of CDK in the initiation step of DNA
replication in eukaryotes. Cell Div. 2, 16
26 Aparicio, O. M., Weinstein, D. M. and Bell, S. P. (1997) Components and dynamics of
DNA replication complexes in S. cerevisiae : redistribution of MCM proteins and Cdc45p
during S phase. Cell 91, 59–69
27 Aparicio, O. M., Stout, A. M. and Bell, S. P. (1999) Differential assembly of Cdc45p and
DNA polymerases at early and late origins of DNA replication. Proc. Natl. Acad. Sci.
U.S.A. 96, 9130–9135
28 Kanemaki, M. and Labib, K. (2006) Distinct roles for Sld3 and GINS during establishment
and progression of eukaryotic DNA replication forks. EMBO J. 25, 1753–1763
29 Moyer, S. E., Lewis, P. W. and Botchan, M. R. (2006) Isolation of the
Cdc45/Mcm2–7/GINS (CMG) complex, a candidate for the eukaryotic DNA replication
fork helicase. Proc. Natl. Acad. Sci. U.S.A. 103, 10236–10241
30 Calzada, A., Hodgson, B., Kanemaki, M., Bueno, A. and Labib, K. (2005) Molecular
anatomy and regulation of a stable replisome at a paused eukaryotic DNA replication fork.
Genes Dev. 19, 1905–1919
31 Kamimura, Y., Masumoto, H., Sugino, A. and Araki, H. (1998) Sld2, which interacts with
Dpb11 in Saccharomyces cerevisiae , is required for chromosomal DNA replication. Mol.
Cell. Biol. 18, 6102–6109
32 Giaever, G., Chu, A. M., Ni, L., Connelly, C., Riles, L., Véronneau, S., Dow, S.,
Lucau-Danila, A., Anderson, K., André, B. et al. (2002) Functional profiling of the
Saccharomyces cerevisiae genome. Nature 418, 387–391
33 Dohmen, R. J., Wu, P. and Varshavsky, A. (1994) Heat-inducible degron: a method for
constructing temperature-sensitive mutants. Science 263, 1273–1276
34 Strahl-Bolsinger, S., Hecht, A., Luo, K. and Grunstein, M. (1997) SIR2 and SIR4
interactions differ in core and extended telomeric heterochromatin in yeast. Genes Dev.
11, 83–93
35 Yabuuchi, H., Yamada, Y., Uchida, T., Sunathvanichkul, T., Nakagawa, T. and Masukata, H.
(2006) Ordered assembly of Sld3, GINS and Cdc45 is distinctly regulated by DDK and
CDK for activation of replication origins. EMBO J. 25, 4663–4674
36 Gomez, E. B., Angeles, V. T. and Forsburg, S. L. (2005) A screen for
Schizosaccharomyces pombe mutants defective in rereplication identifies new alleles of
rad4+ , cut9+ and psf2+ . Genetics 169, 77–89
37 Boe, C. A., Garcia, I., Pai, C. C., Sharom, J. R., Skjolberg, H. C., Boye, E., Kearsey, S.,
MacNeill, S. A., Tyers, M. D. and Grallert, B. (2008) Rapid regulation of protein activity in
fission yeast. BMC Cell Biol. 9, 23
38 Pai, C. C., Garcia, I., Wang, S. W., Cotterill, S., MacNeill, S. A. and Kearsey, S. E. (2009)
GINS inactivation phenotypes reveal two pathways for chromatin association of replicative
α and ε DNA polymerases in fission yeast. Mol. Biol. Cell 20, 1213–1222
39 Picard, D. (2000) Post-translational regulation of proteins by fusions to steroid-binding
domains. Methods Enzymol. 327, 385–401
40 Akman, G. and MacNeill, S. A. (2009) MCM-GINS and MCM-MCM interactions in vivo
visualised by bimolecular fluorescence complementation in fission yeast. BMC Cell Biol.
10, 12
41 Hu, C. D. and Kerppola, T. K. (2003) Simultaneous visualization of multiple protein
interactions in living cells using multicolor fluorescence complementation analysis. Nat.
Biotechnol 21, 539–545
42 Blow, J. J. (2001) Control of chromosomal DNA replication in the early Xenopus embryo.
EMBO J. 20, 3293–3297
43 Boskovic, J., Coloma, J., Aparicio, T., Zhou, M., Robinson, C. V., Mendez, J. and Montoya,
G. (2007) Molecular architecture of the human GINS complex. EMBO Rep. 8, 678–684
Received 5 October 2009/6 November 2009; accepted 20 November 2009
Published on the Internet 15 January 2010, doi:10.1042/BJ20091531
c The Authors Journal compilation c 2010 Biochemical Society
44 De Falco, M., Ferrari, E., De Felice, M., Rossi, M., Hubscher, U. and Pisani, F. M. (2007)
The human GINS complex binds to and specifically stimulates human DNA polymerase
α-primase. EMBO Rep. 8, 99–103
45 Aparicio, T., Guillou, E., Coloma, J., Montoya, G. and Mendez, J. (2009) The human GINS
complex associates with Cdc45 and MCM and is essential for DNA replication. Nucl.
Acids Res. 37, 2087–2095
46 Carthew, R. W. and Sontheimer, E. J. (2009) Origins and mechanisms of miRNAs and
siRNAs. Cell 136, 642–655
47 Grimm, D. (2009) Small silencing RNAs: state-of-the-art. Adv. Drug Deliv. Rev. 61,
672–703
48 Barkley, L. R., Song, I. Y., Zou, Y. and Vaziri, C. (2009) Reduced expression of GINS
complex members induces hallmarks of pre-malignancy in primary untransformed human
cells. Cell Cycle 8, 1577–1588
49 Walter, J. and Newport, J. (2000) Initiation of eukaryotic DNA replication: origin
unwinding and sequential chromatin association of Cdc45, RPA, and DNA polymerase α.
Mol. Cell 5, 617–627
50 Pacek, M., Tutter, A. V., Kubota, Y., Takisawa, H. and Walter, J. C. (2006) Localization of
MCM2–7, Cdc45, and GINS to the site of DNA unwinding during eukaryotic DNA
replication. Mol. Cell 21, 581–587
51 Barry, E. R. and Bell, S. D. (2006) DNA replication in the archaea. Microbiol. Mol. Biol.
Rev. 70, 876–887
52 Brochier-Armanet, C., Boussau, B., Gribaldo, S. and Forterre, P. (2008) Mesophilic
Crenarchaeota : proposal for a third archaeal phylum, the Thaumarchaeota . Nat. Rev.
Microbiol. 6, 245–252
53 Elkins, J. G., Podar, M., Graham, D. E., Makarova, K. S., Wolf, Y., Randau, L., Hedlund,
B. P., Brochier-Armanet, C., Kunin, V., Anderson, I. et al. (2008) A korarchaeal genome
reveals insights into the evolution of the Archaea. Proc. Natl. Acad. Sci. U.S.A. 105,
8102–8107
54 Berthon, J., Cortez, D. and Forterre, P. (2008) Genomic context analysis in Archaea
suggests previously unrecognized links between DNA replication and translation.
Genome Biol. 9, R71
55 Berthon, J., Fujikane, R. and Forterre, P. (2009) When DNA replication and protein
synthesis come together. Trends Biochem. Sci. 34, 429–434
56 Ruchaud, S., Carmena, M. and Earnshaw, W. C. (2007) Chromosomal passengers:
conducting cell division. Nat. Rev. Cell Mol. Biol. 8, 798–812
57 Yoshida, K., Kuo, F., George, E. L., Sharpe, A. H. and Dutta, A. (2001) Requirement of
CDC45 for postimplantation mouse development. Mol. Cell Biol. 21, 4598–4603
58 Pruitt, S. C., Bailey, K. J. and Freeland, A. (2007) Reduced Mcm2 expression results in
severe stem/progenitor cell deficiency and cancer. Stem Cells 25, 3121–3132
59 Shima, N., Alcaraz, A., Liachko, I., Buske, T. R., Andrews, C. A., Munroe, R. J., Hartford,
S. A., Tye, B. K. and Schimenti, J. C. (2007) A viable allele of Mcm4 causes chromosome
instability and mammary adenocarcinomas in mice. Nat. Genet. 39, 93–98
60 Shima, N., Buske, T. R. and Schimenti, J. C. (2007) Genetic screen for chromosome
instability in mice: Mcm4 and breast cancer. Cell Cycle 6, 1135–1140
61 Cimprich, K. A. and Cortez, D. (2008) ATR: an essential regulator of genome integrity. Nat.
Rev. Mol. Cell Biol. 9, 616–627
62 Paulsen, R. D. and Cimprich, K. A. (2007) The ATR pathway: fine-tuning the fork. DNA
Repair 6, 953–966
63 Traven, A. and Heierhorst, J. (2005) SQ/TQ cluster domains: concentrated ATM/ATR
kinase phosphorylation site regions in DNA-damage-response proteins. Bioessays 27,
397–407
64 Han, Y., Ueno, M., Nagahama, Y. and Takakura, N. (2009) Identification and
characterization of stem cell-specific transcription of PSF1 in spermatogenesis. Biochem.
Biophys. Res. Commun. 380, 609–613
65 Takahashi, T. S., Wigley, D. B. and Walter, J. C. (2005) Pumps, paradoxes and
ploughshares: mechanism of the MCM2–7 DNA helicase. Trends Biochem Sci. 30,
437–444
66 MacNeill, S. A. (2009) The haloarchaeal chromosome replication machinery. Biochem.
Soc. Trans. 37, 108–113
Biochem. J. (2010) 425, 489–500 (Printed in Great Britain)
doi:10.1042/BJ20091531
SUPPLEMENTARY ONLINE DATA
Structure and function of the GINS complex, a key component of the
eukaryotic replisome
Stuart A. MACNEILL1
Biomedical Sciences Research Complex, School of Biology, University of St Andrews, North Haugh, St Andrews, Fife KY16 9ST, U.K.
Table S1
Entrez GeneIDs for genes in Figure 2(B) in the main paper
Species
Gins23
MCM
PCNA
Sso
Ape
Pae
Mth
Tac
Mac
Mhu
Hma
Afu
Mma
Pfu
Neq
Csy
Kcr
1455037
1445714
1465397
1455038
1445715
1454088
1444627
1463798
1471029
PriL
PriS
Gins51
1454089
1444626
1463797
1454090
1444625
1463796
1471028
1456561
1472539
3922509
3128974
1484558
2761381
1468847
2654353
6372034
6094809
1472540
3922510
3128913
1468324
1468323
2761382
1468848
6371904
6094628
6371903
6094629
6094806
6094807
6094808
Received 5 October 2009/6 November 2009; accepted 20 November 2009
Published on the Internet 15 January 2010, doi:10.1042/BJ20091531
1
email [email protected]
c The Authors Journal compilation c 2010 Biochemical Society