ArticlePDF AvailableLiterature Review

DNA-damage repair; the good, the bad, and the ugly

Authors:

Abstract

Organisms have developed several DNA-repair pathways as well as DNA-damage checkpoints to cope with the frequent challenge of endogenous and exogenous DNA insults. In the absence or impairment of such repair or checkpoint mechanisms, the genomic integrity of the organism is often compromised. This review will focus on the functional consequences of impaired DNA-repair pathways. Although each pathway is addressed individually, it is essential to note that cross talk exists between repair pathways, and that there are instances in which a DNA-repair protein is involved in more than one pathway. It is also important to integrate DNA-repair process with DNA-damage checkpoints and cell survival, to gain a better understanding of the consequences of compromised DNA repair at both cellular and organismic levels. Functional consequences associated with impaired DNA repair include embryonic lethality, shortened life span, rapid ageing, impaired growth, and a variety of syndromes, including a pronounced manifestation of cancer.
Focus Quality Control
DNA-damage repair; the good, the bad,
and the ugly
Razqallah Hakem
1,2,
*
1
Department of Medical Biophysics, Ontario Cancer Institute/UHN,
University of Toronto, Toronto, Ontario, Canada and
2
Department
of Immunology, Ontario Cancer Institute/UHN, University of Toronto,
Toronto, Ontario, Canada
Organisms have developed several DNA-repair pathways
as well as DNA-damage checkpoints to cope with the
frequent challenge of endogenous and exogenous DNA
insults. In the absence or impairment of such repair or
checkpoint mechanisms, the genomic integrity of the
organism is often compromised. This review will focus
on the functional consequences of impaired DNA-repair
pathways. Although each pathway is addressed individu-
ally, it is essential to note that cross talk exists between
repair pathways, and that there are instances in which a
DNA-repair protein is involved in more than one pathway.
It is also important to integrate DNA-repair process with
DNA-damage checkpoints and cell survival, to gain a
better understanding of the consequences of compromised
DNA repair at both cellular and organismic levels.
Functional consequences associated with impaired DNA
repair include embryonic lethality, shortened life span,
rapid ageing, impaired growth, and a variety of syndromes,
including a pronounced manifestation of cancer.
The EMBO Journal (2008) 27, 589–605. doi:10.1038/
emboj.2008.15
Subject Categories: molecular biology of disease
Keywords: cancer; DNA repair; mouse models; syndromes
Introduction
Organisms have evolved to efficiently respond to DNA
insults that result from either endogenous sources (cellular
metabolic processes) or exogenous sources (environmental
factors). Endogenous sources of DNA damage include
hydrolysis, oxidation, alkylation, and mismatch of
DNA bases; sources for exogenous DNA damage include
ionizing radiation (IR), ultraviolet (UV) radiation, and
various chemicals agents. At the cellular level, damaged
DNA that is not properly repaired can lead to genomic
instability, apoptosis, or senescence, which can greatly
affect the organism’s development and ageing process.
More importantly, loss of genomic integrity predisposes the
organism to immunodeficiency, neurological disorders, and
cancer (O’Driscoll and Jeggo, 2006; Subba Rao, 2007; Thoms
et al, 2007). Therefore, it is essential for cells to efficiently
respond to DNA damage through coordinated and integrated
DNA-damage checkpoints and repair pathways.
DNA-damage checkpoints
The mechanisms of DNA-damage checkpoints are best
understood during their responses to double-strand breaks
(DSBs). Initiation of these checkpoints is dependent on the
transient recruitment of the MRE11/RAD50/NBS1 (MRN)
complex at DSB sites, followed by the recruitment/activation
of ataxia–telangiectasia mutated (ATM) a member of the
family of phosphoinositide-3-kinase-related kinases (PIKKs)
(Su, 2006). In addition, two other PIKKs, DNA-dependent
protein kinase (DNA–PK) and ATR (ATM and Rad3 related),
are also activated and involved in the response to DSBs.
However, the primary function of ATR is the initiation of
DNA-damage response to stalled replication forks (RFs)
(Su, 2006). ATM, ATR, and DNA–PK phosphorylate various
targets that contribute to the overall DNA damage response.
Therefore, within minutes of DSB formation, active ATM
phosphorylates different proteins that are essential for
DNA-damage response and repair. An example includes the
histone H2AX that, following its phosphorylation at the site
of DNA damage by ATM, DNA–PK, or ATR (gH2AX), recruits
other proteins and initiates the chromatin-remodelling
process that is essential for the repair of damaged DNA.
Other proteins recruited to sites of DSBs include
MDC1, 53BP1, and BRCA1, all of which are ATM substrates
and mediators in DNA-damage response. The MRN
complex-mediated resection of DSBs is followed by
single-stranded DNA coating with replication protein A
(RPA), which serves to recruit ATR and its binding partner
ATRIP, and subsequent ATR-dependent phosphorylation of
clapsin, Rad17, BRCA1, and others (Su, 2006).
ATM and ATR are essential for the G1/S, intra-S-phase, and
G2/M DNA-damage checkpoints, and are critical for the
maintenance of genomic integrity. Defects in either ATM or
ATR have been associated with human syndromes. AT M
mutations are associated with the human ataxia–telangiecta-
sia (AT), an autosomal recessive disorder characterized by
cerebellar ataxia, progressive mental retardation, impaired
immune functions, neurological problems, and malignancies
(O’Driscoll and Jeggo, 2006). At the cellular level, AT pheno-
types include chromosomal breakage and IR sensitivity.
Similarly, ATR mutations predispose individuals to Seckel
syndrome, a very rare autosomal recessive human disorder
Received: 31 October 2007; accepted: 16 January 2008
*Corresponding author. Department of Medical Biophysics, Ontario
Cancer Institute/UHN, University of Toronto, 610 University Avenue
PMH, Room 10-622, Toronto, Ontario, Canada M5G 2M9.
Tel.: þ1 416 946 2398/4501; ext: 5661; Fax: þ1 416 946 2984;
E-mail: rhakem@uhnres.utoronto.ca
The EMBO Journal (2008) 27, 589–605 |
&
2008 European Molecular Biology Organization |All Rights Reserved 0261-4189/08
www.embojournal.org
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008
EMBO
THE
EMBO
JOURNAL
THE
EMBO
JOURNAL
589
characterized by growth and mental retardation, as well as
microcephaly (O’Driscoll and Jeggo, 2006). Spontaneous and
IR-induced genomic instability and immunological defects
have also been observed in Seckel syndrome patients. In
contrast to ATM and ATR, no human syndrome has yet been
associated with defective DNA–PK. However, studies of
mouse models have linked mutations of DNA–PK to severe
immunodeficiency (see section Non-homologous end joining
repair pathway).
Activated ATM and ATR mediate the phosphorylation and
subsequent activation of Chk2 and Chk1, respectively; this
process is necessary in the induction of phosphorylation of
CDC25A, marking it for proteosomal degradation (Su, 2006).
The consequential loss of CDC25A results in G1/S arrest, due
to the inefficient loading of CDC45 at the origin of replication.
In addition, activated ATM, ATR, DNA–PK, Chk2, and Chk1
all aid in the phosphorylation and activation of p53, a key
player in DNA-damage checkpoints. Activated p53 transacti-
vates p21, which inhibits two G1/S-promoting cyclin-depen-
dent kinases (CDKs), CDK2 and CDK4. This leads to
sustained G1 arrest, which ultimately hampers the replication
of damaged DNA.
The intra-S-phase checkpoint serves to arrest DNA synth-
esis during S phase of cells with damaged DNA (Su, 2006). In
these cells, CDC25A phosphorylation, mediated by Chk2 or
Chk1, leads to its degradation and the subsequent inactiva-
tion of the S-phase cyclin E/CDK2 complex. Consequently,
these events prevent the loading of CDC45 at the origin of
replication and result in intra-S-phase arrest. It has been
reported that other proteins, including Nijmegen breakage
syndrome 1 (NBS1), BRCA1, SMC1, 53BP1, and MDC1, all
contribute to the intra-S-phase checkpoint.
The activation of the G2/M DNA-damage checkpoint pre-
vents mitotic entry of the damaged cells (Su, 2006). This
checkpoint is mediated by the dual-specificity phosphatase
CDC25C, as well as p53. In normal conditions, CDC25C
dephosphorylates CDC2, allowing the CDC2–cyclin B kinase
to facilitate entry into mitosis. However, phosphorylation of
CDC25C by Chk2 or Chk1, initiates its binding with 14-3-3,
which leads to its cytoplasmic sequestration away from its
substrate, thus preventing mitotic entry. p53 also contributes
to the G2/M checkpoint through its transactivation of p21 and
14-3-3. P21 effectively blocks the phosphorylation of CDC2,
initiating the onset of the G2/M cell-cycle arrest. 14-3-3
sequesters CDC25C in the cytoplasm and promotes the
activation of Wee1, a tyrosine kinase that negatively regulates
CDC2, thus blocking entry into mitosis.
The activation of DNA-damage checkpoints enforces the
growth arrest of damaged cells and allows the DNA-repair
mechanisms to mend the damaged DNA. Once repair
is completed, cells are able to exit the checkpoints and
resume their cell-cycle progression and functions. However,
unsuccessful DNA repair leads to p53-dependent apoptosis
(Chipuk and Green, 2006), in addition to senescence (Collado
et al, 2007).
Defects of DNA-damage checkpoints, similar to impaired
DNA-damage repair, promote genomic instability and predis-
pose individuals to immunodeficiency, neurological defects,
and cancer (Niida and Nakanishi, 2006). Although important
advances have been made in understanding the cellular
mechanisms behind the initiation and maintenance of check-
points, the mechanisms that control checkpoints exit, as well
as how the cell decides survival, death, or senescence,
require further investigation.
Defects associated with DNA-damage
repair pathways
Different DNA-repair pathways exist and perform major roles
at both cellular and organismic levels. These pathways
include (1) the direct reversal pathway, (2) the mismatch
repair (MMR) pathway, (3) the nucleotide excision repair
(NER) pathway, (4) the base excision repair (BER) pathway,
(5) the homologous recombination (HR) pathway, and (6) the
non-homologous end joining (NHEJ) pathway (Figure 1). The
mechanisms for these pathways will not be discussed in
detail in this review; instead we will focus on the functional
consequences associated with their defects.
Direct reversal of DNA damage
In contrast to other DNA-damage repair pathways, direct
reversal of DNA damage is not a multistep process and
does not involve multiple proteins (Sedgwick et al, 2007).
Furthermore, unlike excision repair, direct reversal of DNA
damage does not require the excision of the damaged bases.
An example of a DNA lesion that is repaired by direct reversal
is the O
6
-alkylguanine. Alkylating agents can transfer methyl
or ethyl groups to a guanine, thereby modifying the base and
interfering with its pairing with cytosine during DNA replica-
tion. The cytotoxic and mutagenic O
6
alkyl adduct in DNA is
repaired by direct reversal, which is mediated by the enzyme
Ada in Escherichia coli (E.coli) and the mammalian
O
6
-methylguanine-DNA methyltransferase (MGMT). MGMT,
also known as AGT, removes the DNA adducts by transferring
the alkyl group from the oxygen in the DNA to a cysteine
residue in its active site. This reaction leads to the reversal of
the base damage; however, the alkylation of MGMT leads to
its inactivation and subsequent ubiquitination and proteoso-
mal degradation. MGMT has attracted a great deal of atten-
tion, as certain anticancer chemotherapeutic drugs produce
O
6
-alkylguanine, further supporting its role in modulating the
therapeutic response of tumors to these drugs. Mouse models
for Mgmt inactivation have been generated (Tsuzuki et al,
1996b; Glassner et al, 1999). These mutants were viable and
MMR
Direct
reversal
NER
BER
HR
NHEJ
DNA damage repair pathways
CH3
Figure 1 DNA-repair pathways. Several DNA-repair pathways exist
and deal with various types of DNA insults. These pathways include
(1) the direct reversal pathway, (2) the MMR pathway, (3) the
NER pathway, (4) the BER pathway, (5) the HR pathway, and (6) the
NHEJ pathway.
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization590
showed no increase in spontaneous tumorigenesis (Table I).
However, Mgmt homozygous mice and cells were highly
sensitive to chemotherapeutic alkylating agents such as
methylnitrosourea. Mgmt homozygous mutant females, but
not males, developed larger numbers of dimethylnitrosa-
mine-induced liver and lung tumors compared with controls
(Iwakuma et al, 1997). Additionally, transgenic mice
over-expressing human MGMT or E. coli Ada have also
been generated. In response to alkylating carcinogens that
produce O
6
-alkylguanine in DNA, these transgenic mice
demonstrated a significantly reduced susceptibility to
developing cancers, including thymomas (Dumenco et al,
1993), liver tumors (Nakatsuru et al, 1993), and skin tumors
(Becker et al, 1997).
AlkB is another enzyme that mediates direct DNA damage
reversal in E. coli. This dioxygenase is involved in the repair
of alkylation damage, particularly 1-methyladenine (1meA)
and 3-methylcytosine (3meC). Two mammalian AlkB homo-
logues, ABH2 and ABH3, have been shown to possess DNA-
repair functions similar to the bacterial AlkB (Duncan et al,
2002; Sedgwick et al, 2007). Similar to AlkB, ABH2 and ABH3
have the ability to repair 1meA and 3meC residues. However,
whereas ABH2 prefers double-stranded DNA, ABH3 and AlkB
favour single-stranded DNA and RNA (Aas et al, 2003; Falnes
et al, 2004). Further insight into the function of the mamma-
lian ABH2 and ABH3 came from studies of mice carrying
targeted mutations of these genes. Mice deficient in Abh2,
Abh3, or both, were viable (Ringvoll et al, 2006). Abh2
/
,
but not Abh3
/
, mice showed age-dependent accumulation
of 1meA in their genomic DNA. As in AlkB mutants in E.coli,
mouse embryonic fibroblasts (MEFs) deficient in Abh2
were hypersensitive to methyl methane-sulfonate (MMS)
treatment. However, mice deficient in Abh2 or Abh3 did
not show increased spontaneous cancer development
(Table I). Further studies are required to assess the role
of these dioxygenases, and other AlkB homologues, in
alkylation damage-induced cancer.
These examples of direct DNA-damage reversal mediated
by MGMT/Ada or ABH/AlkB demonstrate the conserved role
of this mechanism in DNA repair. In addition, increased
tumorigenesis of Mgmt mutants, together with the resistance
of MGMT transgenic mice to alkylating carcinogens that
produce O
6
-alkylguanine, further demonstrate the important
role that direct reversal plays in cancer.
The MMR pathway
The MMR pathway plays an important role in both prokar-
yotes and eukaryotes in repairing mismatches, which are
small insertions and deletions that take place during
DNA replication (Figure 1; Jiricny, 2006). Failure of MMR
commonly results in microsatellite instability (MSI). Several
homologues of the bacterial MMR genes MutS and MutL have
been identified in yeast and mammals.
The importance of the MMR pathway became evident
upon identification of mutations in certain human MMR
genes in hereditary non-polyposis colorectal cancer
(HNPCC), a highly penetrant autosomal dominant cancer
syndrome (Figure 2; Vasen et al, 2007). HNPCC, also
known as Lynch syndrome, is characterized by early-onset
colorectal cancer, with elevated levels of MSI in the tumors.
Individuals with HNPCC have an approximate 80% lifetime
risk for colorectal cancer, and are also predisposed to the
development of endometrial, ovarian, gastric, and other types
of malignancies.
Approximately 70–80% of germline mutations identified in
HNPCC families are mutations in MLH1 or MSH2, whereas
mutations in MSH6 are found in approximately 10% of
HNPCC families (Peltomaki and Vasen, 1997). Germline
mutations in other human MMR genes, including PMS1,
PMS2,MLH3, and exonuclease 1 (EXO1), have also been
found in HNPCC families; however, they occur at a much
lower frequency (Vasen et al, 2007). In addition, inactivation
of MLH1 by mutations at the promoter or coding sequences,
or by promoter methylation, has been identified in sporadic
colorectal tumors (Kane et al, 1997; Veigl et al, 1998). Recent
Table I Examples of mouse models for direct reversal
Genotype Developmental
defects
Fertility defects Spontaneous
tumorigenesis
Induced tumorigenesis References
Mgmt/None None Not affected Increased DMNA-
induced lung and
liver cancer in females
(Tsuzuki et al, 1996a, b;
Iwakuma et al, 1997;
Glassner et al, 1999)
Abh2/
Abh3/
Abh2/Abh3/
None None Not affected Not tested (Ringvoll et al, 2006)
Defective repair
HR pathway NHEJ pathway
–SCID
–RS-SCID
BER pathway
–MAP
MMR pathway
–HNPCC
NER pathway
–XP
–CS
–TTD
–BS
–WS
–RTS
–ATLD
–NBS
Figure 2 Examples of human syndromes and disorders associated
with defective DNA-damage repair. Impaired MMR pathway leads to
the hereditary HNPCC. Mutations of certain human NER genes have
been associated with syndromes and disorders including the XP, CS,
and TTD. MAP, a rare disorder, has been shown associated with
mutations of the BER gene MUTYH. Various human syndromes and
disorders have been associated with defects of the HR pathway.
They include ATLD, NBS, BS, WS and RTS. Mutations of certain
human genes involved in NHEJ lead to the SCID or RS-SCID.
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 591
studies, although very limited, have identified rare patients
with homozygous germline mutations for MLH1,MSH2,
MSH6,orPMS2 (Felton et al, 2007). Typically, these indivi-
duals have a reduced life span and, in contrast to hetero-
zygous MMR individuals, tend to develop juvenile
haematological malignancies and brain cancer.
In yeast, Msh2 forms heterodimers with Msh3 and Msh6,
proteins that bind DNA mismatches and initiate the MMR
process. In Saccharomyces cerevisiae (S.cerevisiae), Msh2,
Msh3, and Msh6 mutants are viable (Marsischky et al, 1996).
Both Msh2 and Msh6 S.cerevisiae mutants show high fre-
quencies of base substitution, whereas only Msh2 mutants
exhibit high frameshift mutations. Msh3 mutations in
S.cerevisiae result in low rates of frameshift mutations.
However, on Msh6-mutant background, synergistic effects
of the dual mutations have been observed, including in-
creased MSI and mutability similar to Msh2 mutants.
Mutant mice for MutS and MutL MMR homologues have
also been generated using gene targeting (Table II). Mutant
mice for the MutS homologues include Msh2,Msh3,Msh4,
Msh5, and Msh6. Mice carrying homozygous mutations for
Msh4 or Msh5 did not exhibit cancer phenotypes; however,
males and females were infertile, consistent with the role of
MutS homologues in processing meiotic recombination inter-
mediates (de Vries et al, 1999; Edelmann et al, 1999). In
contrast, homozygous mutants for Msh2 (de Wind et al,
1995; Reitmair et al, 1995), Msh3 (Edelmann et al, 2000),
and Msh6 (Edelmann et al, 1997) have increased risk for
developing cancers such as lymphoma, gastrointestinal, and
skin cancer.
Mutant mice for MutL homologues include Pms1
/
,
Pms2
/
,Mlh1
/
, and Mlh3
/
mice (Table II). These
mutants are viable; however, males and females deficient in
Mlh1 (Baker et al, 1996; Edelmann et al, 1996) or Mlh3
(Lipkin et al, 2002) and males deficient in Pms2 (Baker
et al, 1995) are sterile, demonstrating a requirement for
these proteins during meiosis. In addition, mouse MutL
homologues are differentially required for cancer suppres-
sion. Pms1
/
mice do not show any increased risk for cancer
(Prolla et al, 1998), whereas Mlh1
/
(Prolla et al, 1998; Chen
et al, 2005) and Mlh3
/
mice (Chen et al, 2005) are predis-
posed to developing lymphomas and gastrointestinal tumors.
Similarly, Pms2-null mutants (Prolla et al, 1998; Chen et al,
2005) are prone for lymphoma development.
EXOI physically interacts with MSH2, MSH3, and MLH1,
and is involved in the excision of mismatched bases in DNA
(Tishkoff et al, 1997; Schmutte et al, 2001). Mutant mice for
ExoI have impaired MMR, accumulate MSI, and exhibit a
greater risk for developing lymphomas (Wei et al, 2003).
These mutants also have meiotic defects and are sterile,
demonstrating the requirement of ExoI in meiosis.
Double mutant mice carrying dual mutations of different
MMR genes have also been reported. For example, Msh3-
mutant mice develop cancer with low frequency and at a later
age, whereas Msh3
/
Msh6
/
mice (Edelmann et al, 2000)
die prematurely and develop tumors including lymphomas,
gastrointestinal, and skin tumors. This phenotypic outcome
is similar to that of Msh2
/
or Mlh1
/
mice that are the
most cancer-prone MMR mutants, as half of these mutants
die around 6 months of age. This cooperation between
mutations of Msh2 and Msh6 in mice is reminiscent of their
collaboration in the maintenance of genomic integrity of
S.cerevisiae.
Immunoglobulin (Ig) diversification, an essential process
for immunity, involves somatic hypermutation (SHM) of the
Ig genes, as well as VDJ recombination and class-switch
recombination (CSR), two processes mediated by NHEJ
(Maizels, 2005). Interestingly, studies of the various MMR-mutant
strains have implicated a role for this pathway in SHM and CSR.
Thus, Msh2,Msh6,andExoI, but not Msh3-mutant mice, have
reduced CSR and SHM (Rada et al, 1998; Wiesendanger et al,
2000; Bardwell et al, 2004; Li et al, 2004).
Table II Examples of mouse models for the MMR pathway
Genotype Developmental
defects
Fertility defects Spontaneous tumorigenesis References
Msh2/None None High frequency and early onset
of lymphomas, gastrointestinal,
and skin cancer
(de Wind et al, 1995;
Reitmair et al, 1995)
Msh3/None None Low frequency and late onset
of lymphomas, gastrointestinal,
and skin cancer
(Edelmann et al, 2000)
Msh4/None Infertile Not affected (Kneitz et al, 2000)
Msh5/None Infertile Not affected (de Vries et al, 1999;
Edelmann
et al, 1999)
Msh6/None Unaffected Lymphoma, gastrointestinal, and
skin cancer
(Edelmann et al, 1997)
Msh3/Msh6/None None Higher frequency of lymphomas,
gastrointestinal, and skin
tumours compared to single
mutants
(Edelmann et al, 2000)
Pms1/None None Not affected (Prolla et al, 1998)
Pms2/None Male infertility Lymphomas (Baker et al, 1995; Prolla et al,
1998; Chen et al, 2005)
Mlh1/None Infertile High frequency and early onset
of lymphomas and gastrointest-
inal tumours
(Baker et al, 1996; Edelmann
et al, 1996; Prolla et al, 1998;
Chen et al, 2005)
Mlh3/None Infertile Lymphomas and gastrointestinal
tumours
(Lipkin et al, 2002; Chen et al,
2005)
ExoI/None Infertile Lymphomas (Wei et al, 2003)
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization592
Whereas most HNPCC human individuals carry heterozy-
gous germline mutations of MMR genes, which predisposes
them to cancer, mice heterozygous for MMR mutations do
not appear to have an increased risk for developing cancer.
This difference is not specific for MMR mutations, as
heterozygous mutations in certain genes involved in other
DNA-damage repair pathways are also able to predispose
humans, but not mice, to cancer. The reasons for these
differences remain unknown, although species differences
in the DNA-damage repair pathways, metabolism, or life
span could contribute to these observed human–mouse
discrepancies. Despite these differences, mouse models
have significantly improved our understanding of the MMR
and other repair mechanisms, and their roles in preserving
genomic integrity and suppressing cancer.
The NER pathway
The NER pathway is a multistep process that serves to repair
a variety of DNA damage, including DNA lesions caused by
UV radiation, mutagenic chemicals, or chemotherapeutic
drugs that form bulky DNA adducts (Figure 1; Leibeling
et al, 2006). Over 30 different proteins are involved in the
mammalian NER, whereas only three proteins (UvrA, UvrB,
and UvrC) are required by prokaryotes (Truglio et al, 2006).
Two NER sub-pathways that have been identified are as
follows: the global genome NER (GG-NER) that detects and
removes lesions throughout the genome, and the transcrip-
tion-coupled NER (TC-NER), which repairs actively tran-
scribed genes. NER begins with the recognition of the DNA
lesion, followed by incisions at sites flanking the DNA lesion,
and culminates in the removal of the oligonucleotide contain-
ing the DNA lesion. Ligation of a newly synthesized oligonu-
cleotide, complementary to the pre-existing strand, serves to
fill the gap, thus ending the NER process. GG-NER and
TC-NER involve several common proteins and proceed
through the same repair steps, except during recognition of
the DNA lesion. In GG-NER this recognition involves the
XPC–RAD23B and DDB1–DDB2/XPE proteins, whereas re-
cognition in TC-NER is mediated by cockayne syndrome
group A (CSA) (ERCC8) and CSB (ERCC6). NER has attracted
a great deal of attention due to its role in three rare human
syndromes characterized by increased cancer frequencies,
neurodegeneration and ageing (Figure 2). These syndromes
are xeroderma pigmentosum (XP), Cockayne syndrome (CS),
and trichothiodystrophy (TTD) (Thoms et al, 2007).
XP individuals show extremely severe skin sensitivity to
short intervals of sun exposure and most develop freckles at
an early age. In addition, XP individuals may exhibit eye
damage as they suffer chronic UV-induced conjunctivitis and
keratitis as a consequence of continual sun exposure. XP
individuals have greater than 1000-fold increased skin cancer
risk, which first appears at an average age of 10 years.
Approximately 20% of XP individuals also develop neurolo-
gical abnormalities. XP is caused by mutations of the NER
gene XPA,XPB (ERCC3), XPC,XPD (ERCC2), XPE (DDB2), or
XPF; XPG. Whereas XP individuals carrying mutations of XPC
or XPE (DDB2) are only deficient in GG-NER, the remaining
XP individuals are deficient in both GG-NER and TC-NER
(Thoms et al, 2007).
CS is a very rare human autosomal recessive inherited
genetic disease (Thoms et al, 2007). Similar to XP individuals,
CS individuals suffer excessive sun sensitivity, but without
increased predisposition for skin cancer. Common CS symp-
toms include growth retardation (dwarfism), progressive
cognitive impairment, and ophthalmologic disorders such
as cataracts or retinitis pigmentosa. CS individuals typically
die in the first or second decade of life. CS is caused by
mutations of either CSA or CSB, two proteins essential
for DNA-damage recognition and initiation of TC-NER.
Therefore, although CS individuals are deficient in TC-NER,
they remain proficient in GG-NER.
TTD is a rare human autosomal recessive disorder asso-
ciated with defective NER; the most severe cases are asso-
ciated with mutations in the XPB or XPD genes. Clinical
characteristics of TTD include brittle hair and nails, dwarf-
ism, and ataxia (Thoms et al, 2007). In addition, half of TTD
individuals exhibit sensitivity to sunlight; however, skin
cancer predisposition has not been linked to this syndrome.
Several mouse models for NER mutations have been
generated (Table III). Homozygous mutants for XP genes
are viable (Nakane et al, 1995; Sands et al, 1995; Harada
et al, 1999; Itoh et al, 2004; Tian et al, 2004a, b; Yoon et al,
2005), with the exception of the pre-implantation embryonic
lethality of Xpd mutants (de Boer et al, 1998b). Xpd
R722W
-
mutant mice carrying an amino-acid substitution that mimics
a human XPD allele associated with TTD have been gener-
ated (de Boer et al, 1998a). Similarly, Xpd
G602D
-mutant mice
carrying a substitution at amino acid 602 have been gener-
ated to mimic the human combined XP/CS (Andressoo et al,
2006). Both Xpd
R722W
and Xpd
G602D
mutants have been
proven viable and reproduced some of the characteristics of
individuals that carry these XPD mutations.
With the exception of Xpe (Ddb2) mutants (Itoh et al,
2004; Yoon et al, 2005), homozygous mutant mouse cells for
Xpa (Nakane et al, 1995), Xpc (Sands et al, 1995), Xpf (Tian
et al, 2004b), Xpg (Harada et al, 1999; Tian et al, 2004a),
Xpd
R722W
(de Boer et al, 1998a), or Xpd
G602D
(Andressoo
et al, 2006) were all UV sensitive, correlating to the results
obtained from human mutant cells. Increased predisposition
for UV-induced skin cancer was observed with Xpa,Xpc,
Xpd
R722W
,Xpd
G602D
, and Xpe (Ddb2)-mutant mice (Nakane
et al, 1995; Sands et al, 1995; de Boer et al, 1999; Itoh et al,
2004; Yoon et al, 2005; Andressoo et al, 2006). Thus, as seen
in humans, XP proteins play a major role in murine NER.
DDB1 and HR23B are two proteins involved with XPC in
the recognition of DNA damage and the initiation of GG-NER.
However, in contrast to Xpc
/
mice (Sands et al, 1995),
Ddb1 mutants die during early embryonic development
(Cang et al, 2006). Inactivation of Ddb1 in developing CNS
and lens resulted in massive p53-dependent apoptosis of
dividing cells and lethality just after birth (Cang et al,
2006). MEFs deficient in Ddb1 showed defective proliferation
and were UV-sensitive. A total of 90% of HR23B mutants
suffer intrauterine or neonatal death (Ng et al, 2002). The
surviving HR23B
/
mice were growth retarded and males
were sterile; however, their NER and UV sensitivity remained
normal. In contrast, mutants for HR23A, another homologue
of the S.cerevisiae NER gene Rad23, are viable and their cells
proficient in UV responses (Ng et al,2003).However,mutations
of both HR23A and HR23B lead to embryonic lethality and
increased UV sensitivity, showing a redundancy between these
two NER proteins (Ng et al, 2003). It remains to be shown
whether inactivation of Ddb1 or mHR23A/B would facilitate
spontaneous or UV-induced cancer in mouse models.
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 593
Mice mutant for Csa (Ercc8)orCsb (Ercc6) have been
generated (van der Horst et al, 1997, 2002). These mutants
are viable, but their cells, although competent in GG-NER, are
nonetheless impaired in TC-NER and are UV-sensitive.
Surprisingly, in contrast to CS individuals, Csa- and Csb-
mutant mice are prone to UV-induced skin cancer. This
discrepancy is also evident in TTD mouse models, as in
contrast to TTD individuals, Xpd
R722W
mice are susceptible
to UV-induced skin cancer (de Boer et al, 1999).
As mentioned earlier, cross talk exists between DNA-
damage repair pathways and certain DNA-damage repair
proteins. For example, during NER, the endonuclease
ERCC1/XPF cleaves a DNA strand on the 50-side of the lesion,
allowing the repair to proceed. However, ERCC1/XPF is also
implicated in interstrand crosslink (ICL) repair and HR,
suggesting that the phenotypes observed in Ercc1 or Xpf
mutants are unlikely to result only from the impairment of
NER. Cells deficient in Ercc1 or Xpf show high sensitivity
to UV and to the ICL mitomycin C (MMC), and mice
with inactivation of Ercc1 or Xpf are runted, dying at
approximately 3–4 weeks of age (McWhir et al, 1993;
Weeda et al, 1997; Tian et al, 2004b).
Thus, NER is an important DNA-repair pathway and its
impairment is associated with growth defects, excessive UV
sensitivity, and in certain cases, increased skin cancer.
The BER pathway
The BER pathway deals with base damage, the most common
insult to cellular DNA (Figure 1; Wilson and Bohr, 2007). Two
sub-pathways, short-patch BER and long-patch BER, are
involved in BER. The short-patch BER sub-pathway typically
replaces a single nucleotide, whereas the long-patch
sub-pathway results in the incorporation of 2–13 nucleotides.
The two sub-pathways progress through different major
processes that initially involve the removal of the damaged
base by glycosylases such as Ogg1 and Mutyh (Myh). This is
followed by strand incision of the apurinic or apurimidinic
(AP) site by the endonuclease APE1. The newly generated
gap is filled by incorporation of nucleotide(s), mediated by
DNA polymerase-b(Polb) in the case of the short-patch BER
sub-pathway, and by Polband/or Poleor din the case of the
long-patch BER sub-pathway. Strand ligation is carried out by
the XRCC1/ligase III (LigIII) complex in the case of the short-
patch sub-pathway. For the long-patch BER sub-pathway,
other proteins are involved in the repair process before the
ligation takes place. Thus, FEN1, PARP1, and LigI participate
in the DNA synthesis–ligation step and displace the 2- to
13-base DNA flap. FEN1, a 50-flap endonuclease and
50–30exonuclease, excises the flap and the strand ligation is
carried out by LigI.
For a period of time, little evidence existed to support the
involvement of BER in human cancer or any other disorders.
However, recent studies have demonstrated the existence of a
human disorder linked to defective BER (Figure 2). This
autosomal recessive disorder, referred to as MUTYH-asso-
ciated polyposis (MAP), is associated with biallelic germline
mutations of the human MUTYH, and is characterized by
multiple colorectal adenomas and carcinomas (Cheadle and
Sampson, 2007). The glycosylase MUTYH, a bacterial mutY
homologue, functions in the BER of oxidative DNA damage
by excising adenines misincorporated opposite 8-oxoG.
Deficient repair of this damage results in G:C-T:A
mutations, typically found in the adenomatous polyposis
coli (APC) gene in MAP tumors.
In addition to Myh, several mammalian glycosylases
are involved in BER, including Nth1, Ogg1, Ung, and Aag.
Table III Examples of mouse models for the NER pathway
Genotype Developmental defects Fertility
defects
Induced tumorigenesis References
Xpd/Pre-implantation
embryonic lethality
NA NA (de Boer et al, 1998b)
Xpd
R722W/R722W
Growth retardation None UV- and DMBA-induced
skin cancer
(de Boer et al, 1998a, 1999)
Xpd
G602D/G602D
Growth retardation None Early onset of UV-induced
skin and/or eye tumours
(Andressoo et al, 2006)
Xpe/None None UV-induced skin cancer (Itoh et al, 2004;
Yoon et al, 2005)
Xpa/None None UV- and DMBA-induced
skin cancer
(Nakane et al, 1995)
Xpc/None None UV-induced skin cancer (Sands et al, 1995)
Xpg/Growth retardation and
premature death
NA NA (Harada et al, 1999)
Ddb1/Early embryonic lethality NA NA (Cang et al, 2006)
HR23A/Unaffected None NT (Ng et al, 2002)
HR23B/Intrauterine/neonatal
death. 10% viable but
growth retarded
Male infertility NT (Ng et al, 2002)
HR23A/;
HR23B/
Embryonic lethality NA NA (Ng et al, 2003)
Csa/None None UV-induced skin cancer (van der Horst et al, 1997)
Csb/None None UV-induced skin cancer (van der Horst et al, 2002)
Ercc1/Growth retardation and
death before weaning
NA NA (McWhir et al, 1993;
Weeda et al, 1997)
Xpf/Growth retardation and
death before weaning
NA NA (Tian et al, 2004b)
NA, not applicable; NT, not tested.
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization594
Murine homozygous mutants for the glycosylases Nth1,
Ogg1,Ung,Aag,andMutyh have been generated by gene
targeting, and surprisingly, these mutants were viable
(Table IV). Nth1
/
,Ogg1
/
, and Aag
/
mice showed no
overt abnormalities (Engelward et al, 1997; Klungland et al,
1999; Ocampo et al, 2002; Takao et al, 2002); however after a
long latency, Ung
/
and Mutyh
/
mice developed B-cell
lymphomas and intestinal tumors, respectively (Nilsen et al,
2003; Sakamoto et al, 2007). These findings suggest func-
tional redundancy of certain BER glycosylases. This is further
supported by the pronounced phenotypes of Ogg1
/
Mutyh
/
mice; these mice had shorter life span and elevated
risk for cancer, with 50% of double mutants developing lung
tumors, lymphomas, sarcomas, and others tumors by
15 months of age (Xie et al, 2004).
In contrast to glycosylases, targeted inactivation of en-
zymes that act downstream of the glycosylases in BER
resulted in embryonic or post-natal lethality. Thus, homo-
zygous mutants for Ape1 (Ludwig et al, 1998), LigI (Petrini
et al, 1995), LigIII (Puebla-Osorio et al, 2006), Xrcc1 (Tebbs
et al, 1999), or Flap endonuclease 1 (Fen1) (Kucherlapati et al,
2002) died during embryonic development, whereas homo-
zygous mutants for Polbdied immediately after birth
(Gu et al, 1994; Sugo et al, 2000). The death of mutants
such as Xrcc1
/
,LigIII
/
and Polb
/
was preceded by
elevated levels of apoptosis (Gu et al, 1994; Tebbs et al,
1999; Sugo et al, 2000; Puebla-Osorio et al, 2006).
Interestingly, inactivation of p53 rescued the apoptosis but
not the lethality of these mutants, suggesting the contribution
of other p53-independent mechanisms in the deaths of
these mutants.
Despite the presumed important role for the BER pathway
in maintaining genomic integrity, mutations in this pathway
have not significantly predisposed mutant mice for cancer.
This is in contrast to mutations of other excision repair
pathways, such as NER and MMR.
The HR repair pathway
DSB repair can be mediated by two major repair pathways
depending on the context of the DNA damage, HR or NHEJ
repair pathways (Figure 1; Kanaar et al, 2008). In bacteria
and yeast, DSBs are preferentially repaired by HR, whereas
more than 90% of DSB in mammalian cells are repaired
by NHEJ. Both pathways are well defined and their
impairment is associated with defects and pathologies,
including increased cell death, cell-cycle arrest, telomere
defects, genomic instability, meiotic defects, immuno-
deficiency, and cancer (Krogh and Symington, 2004; Sung
and Klein, 2006).
HR is a multistep process that requires several proteins and
operates at the S or G2 phase of the cell cycle. Although it
accounts only for the repair of B10% of DSBs in mammalian
cells, HR defects can have severe consequences, as demon-
strated by the human syndromes AT-like disorder (ATLD) and
the NBS (Figure 2; Thompson and Schild, 2002). The predis-
position to either syndrome has been linked to mutations in
the MRN complex, which is important for the resection of
DSBs (Thoms et al, 2007). However, it is important to note
MRN functions are not restricted to HR, as is also involved in
NHEJ, checkpoint activation, and telomere maintenance
(Niida and Nakanishi, 2006).
The very rare human ATLD is associated with hypo-
morphic mutations in the MRE11 gene (Taylor et al, 2004).
Patients with this disorder exhibit clinical features similar to
AT, including immunodeficiency and progressive neurolo-
gical degeneration (Thoms et al, 2007). However, in contrast
to AT, ATLD is not associated with ocular telangiectasia and
the course of the disease is considerably milder. Similar to AT
and NBS, cellular features of ATLD include defective DSB
repair and repair-related cell responses, hypersensitivity to
IR, as well as increased spontaneous and IR-induced genomic
instability.
The NBS is a rare human autosomal recessive disorder
caused by hypomorphic NBS1 (NIBRIN) mutations (Thoms
et al, 2007). This disorder is characterized by growth retarda-
tion, immunodeficiency, microcephaly, and cancer predispo-
sitions, particularly lymphomas. Cellular characteristics of
NBS include radiosensitivity, increased levels of spontaneous
and IR-induced chromosome breakage, and defective
cell-cycle checkpoints.
Table IV Examples of mouse models for the BER pathway
Genotype Developmental
defects
Fertility defects Spontaneous tumorigenesis References
Nth1/None None Not affected (Ocampo et al, 2002;
Ta k a o et al, 2002)
Ogg1/None None Not affected (Klungland et al, 1999)
Ung/None None Late onset of B-cell
lymphomas
(Nilsen et al, 2003)
Aag/None None Not affected (Engelward et al, 1997)
Mutyh/None None Late onset of intestinal
tumours
(Sakamoto et al, 2007)
Ogg1/
Mutyh/
None None Late onset of lung tumours,
lymphomas, sarcomas, and
others tumours
(Xie et al, 2004)
Ape1/Embryonic lethality NA NA (Ludwig et al, 1998)
LigI/Embryonic lethality NA NA (Petrini et al, 1995)
LigIII/Embryonic lethality NA NA (Puebla-Osorio et al, 2006)
Xrcc1/Embryonic lethality NA NA (Tebbs et al, 1999)
Fen1/Embryonic lethality NA NA (Kucherlapati et al, 2002)
Polb/Death immediately after birth NA NA (Gu et al, 1994; Sugo et al,
2000)
NA, not applicable.
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 595
In contrast to both MRE11 and NBS1, no data have yet been
reported to support a role for RAD50 in human chromosomal
breakage or immunodeficiency syndromes.
The functions of the MRN complex have been studied in
various organisms. S.cerevisiae strains carrying null muta-
tions in components of the Mre11p–Rad50p–Xrs2p (MRX)
complex have been generated, where Xrs2p is the functional
homologue of mammalian NBS1. MRX mutants are viable,
hypersensitive to IR and MMS, and are defective in meiotic
recombination (Krogh and Symington, 2004). Mutations of
the MRN complex have also been assessed in DT40. This
chicken B-lymphocyte cell line is deficient in p53, but highly
competent for homologous recombination and gene target-
ing. Deficiency of Mre11 in DT40 is lethal and its conditional
inactivation leads to radiosensitivity, defective proliferation,
genomic instability, and decreased HR (Yamaguchi-Iwai et al,
1999). DT40 deficient in NBS1 also display increased IR
sensitivity, abnormal S-phase checkpoints, and decreased
HR (Tauchi et al, 2002).
In contrast to S.cerevisiae, null mutations of Mre11
(Xiao and Weaver, 1997), Nbs1 (Zhu et al, 2001), and
Rad50 (Luo et al, 1999) result in early mouse embryonic
lethality (Table V). Therefore, to circumvent embryonic leth-
ality, hypomorphic and tissue-specific mutants for the three
MRN components were generated. Mre11
ATLD1/ATLD1
-mutant
mice carrying an A to T substitution at amino acid 1894 of
Mre11, which results in a 75-amino-acid truncation, exhibited
reduced female fertility, defective ATM functions, and
increased genomic instability. However, no cancer pheno-
types were observed (Theunissen et al, 2003).
Rad50
s
(Rad50
k22M
) hypomorphic mutant mice have also
been generated (Bender et al, 2002). Approximately 40% of
Rad50
s/s
mutants die in utero; however, those that are viable
show progressive haematopoietic failure, short life span, and
increased predisposition to thymic lymphoma. In contrast to
rad50
s
in yeast, Rad50
s/s
did not impair meiotic progression
and the mutants were fertile; however, p53-mediated
apoptosis was increased in the testes. Inactivation of the
Table V Examples of mouse models for the HR pathway
Genotype Developmental defects Fertility
defects
Spontaneous
tumorigenesis
References
Mre11/Early embryonic
lethality
NA NA (Xiao and Weaver, 1997)
Mre11
ATLD1/ATLD1
None Reduced female
fertility
None (Theunissen et al, 2003)
Rad50/Early embryonic
lethality
NA NA (Luo et al, 1999)
Rad50
s/s
(Rad50
k22M/k22M
)
40% die in utero. The
one that survive show
haematopoietic failure
None Thymic lymphoma (Bender et al, 2002)
Nbs1/Early embryonic
lethality
NA Mild cancer predisposition
of Nbs1+/mice
(Zhu et al, 2001)
Nbs1
D23/D23
Growth retardation,
lymphoid defects
Female sterility Thymic lymphomas (Kang et al, 2002)
Nbs1
D45/D45
None None Twofold increase (Williams et al, 2002)
Rad52/None None Not affected (Rijkers et al, 1998)
Rad51/,Rad51B/
Rad51D/
Embryonic lethality NA NA (Lim and Hasty, 1996; Tsuzuki
et al, 1996a; Shu et al, 1999;
Pittman and Schimenti, 2000)
Xrcc2/Embryonic lethality NA NA (Deans et al, 2003; Orii et al,
2006; Adam et al, 2007)
Dmc1/None Mutants are
sterile
Not affected (Pittman et al, 1998; Yoshida
et al, 1998)
Rad54/,
Rad54B/, and
Rad54/;Rad54B/
None None Not affected (Essers et al, 1997; Bross et al,
2003; Wesoly et al, 2006)
Brca1/Embryonic lethality NA NA (Gowen et al, 1996; Hakem
et al, 1996; Liu et al, 1996;
Ludwig et al, 1997; Xu et al,
2001)
Brca1 mutation in
mammary epithelial
cells
NA Mammary tumorigenesis
that is enhanced on
Chk2-orp53-mutant
backgrounds
(Xu et al, 1999; McPherson
et al, 2004b)
Brca2/Early embryonic
lethality
NA NA (Ludwig et al, 1997; Suzuki
et al, 1997)
Brca2 mutation in
mammary epithelial
cells
Mammary tumorigenesis
on p53-mutant background
(Jonkers et al, 2001)
Blm/None None Predisposition to a wide
range of tumours
(Luo et al, 2000)
Mus81/None None T- and B-cell lymphomas (McPherson et al, 2004a)
Mus81/p53/Female embryonic
lethality
None Multiples tumours
including lymphomas
and sarcomas
(Pamidi et al, 2007)
NA, not applicable.
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization596
Atm–Chk2–p53 pathway by the hypomorphic mutation
Mre11
ATLD1
or the Nbs1
DB
(Nbs1
D45
) mutation, was able to
rescue the depletion of haematopoeitic cells in the Rad50
s/s
mutants (Morales et al, 2005). Surprisingly, tumorigenesis,
senescence, and radiosensitivity associated with Atm muta-
tion were all partially suppressed by the Rad50
s
mutation
(Morales et al, 2005). Although the exact mechanism for this
rescue is not fully understood, it might involve compensatory
activation of other checkpoint pathways, such as the ATR
pathway.
Similar to Rad50 and Mre11 mutations, Nbs1-null muta-
tions resulted in early embryonic lethality (Zhu et al,2001;
Dumon-Jones et al, 2003). In contrast, Nbs1
D23
and
Nbs1
D45
hypomorphic mutants were viable (Kang et al,
2002; Williams et al, 2002). Cells from these mutants were
defective in the intra-S-phase and G2/M checkpoints.
Heterozygous mice carrying a null Nbs1 mutation (Dumon-
Jones et al, 2003), as well as homozygous mice carrying
Nbs1
D23
or Nbs1
D45
hypomorphic mutations (Kang et al,
2002; Williams et al, 2002), demonstrated a mild pre-
disposition for cancer. Interestingly, whereas p53 inactivation
shortened the tumor latency of Nbs1
D45
mutants, Atm
inactivation, or its impaired function on Mre11
ATLD1/ATLD1
-
mutant background, resulted in synthetic embryonic
lethality of Nbs1
D45
mutants (Williams et al, 2002;
Morales et al, 2005).
In addition to the previous models, a conditional mutant
strain for Nbs1 has also been generated (Frappart et al, 2005).
Neuronal inactivation of Nbs1 in these mice leads to chro-
mosomal breaks, microcephaly, growth retardation, cerebel-
lar defects, and ataxia, representing a combination of features
characteristic of NBS, AT, and ATLD. p53 inactivation in this
model also significantly rescued the neurological defects
associated with Nbs1 mutations (Frappart et al, 2005).
Thus, the MRN complex is essential for maintaining geno-
mic integrity, cell viability, and checkpoint activation.
Moreso, its requirement in various species demonstrates its
essential conserved functions.
Similar to Rad50, Rad52 is a member of the Rad52 epistasis
group of proteins originally identified by their requirement
for the repair of IR-induced DNA damage (Krogh and
Symington, 2004). Inactivation of rad52 in S.cerevisiae
results in increased radiation sensitivity and decreased
recombination (Symington, 2002). However, its inactivation
in DT40 reduced gene-targeting frequency, but did not lead to
increased IR sensitivity and viability and growth of the
mutant cells were not affected (Yamaguchi-Iwai et al,
1998). The effects of inactivation of mouse Rad52 were also
investigated. Null Rad52 mouse embryonic stem (ES) cells
obtained by gene targeting demonstrated a 30–40% decrease
in the frequency of HR compared with controls (Rijkers
et al, 1998). In contrast to rad52 mutants in S.cerevisiae,
Rad52-deficient ES cells were not hypersensitive to IR or
agents that induce DSBs. Rad52
/
mice, were viable, fertile,
and showed no overt abnormalities (Rijkers et al, 1998).
Rad52 inactivation was shown to extend the life span of
Atm
/
mice, partially rescue their T-cell development, and
suppress/delay their tumorigenesis (Treuner et al, 2004).
However, growth defects, infertility, and radiosensitivity of
Atm
/
mice were not rescued on by a Rad52-mutant
background. The reasons for the rescue of certain Atm
/
phenotypes by Rad52 inactivation are not clear, but
nevertheless this is reminiscent of the rescue mediated
by Rad50
s/s
mutation of tumorigenesis, senescence, and
radiosensitivity associated with Atm inactivation (Morales
et al, 2005).
The mammalian Rad51 family is also important for HR.
This family is composed of RAD51, disrupted meiotic cDNA 1
(Dmc1), and five RAD51 paralogues, RAD51B, RAD51C,
RAD51D, XRCC2, and XRCC3. RAD51, a homologue of the
bacterial DNA-dependent ATPase, RecA, is central for the
homology search and strand exchange during HR.
Mutants rad51 in S.cerevisiae are viable, IR sensitive, have
meiotic defects, and accumulate meiosis-specific double-
strand breaks (Shinohara et al, 1992). Rad51 deficiency in
DT40 leads to chromosomal breakage, G2/M arrest, and
cell death (Sonoda et al, 1998). Rad51-null mutation in
mice results in early embryonic lethality associated
with chromosomal loss, radiosensitivity, decreased cell
proliferation, and increased apoptosis (Lim and Hasty,
1996; Tsuzuki et al, 1996a).
Mutants for Rad51 paralogues were also generated. Thus,
for example, inactivation of Rad51 paralogues in DT40 did not
compromise cell survival (Takata et al, 2000, 2001). However,
decreased gene-targeting efficiency, increased spontaneous,
and induced cell death in response to IR or MMC, as well as
elevated chromosomal aberrations, were observed in these
mutants. Mutant mice for the Rad51 paralogue Rad51B (Shu
et al, 1999), Rad51D (Pittman and Schimenti, 2000), or Xrcc2
(Deans et al, 2003) were also generated. These mutations
resulted in lethality at different stages of embryonic develop-
ment and their effects on cell growth were variable.
Consistent with the role of Rad51 and its paralogues in HR,
their mutations result in accumulation of DNA damage
and activation of cellular checkpoints, including p53.
Consequently, the embryonic lethality of Rad51 (Lim and
Hasty, 1996), Rad51 B (Shu et al, 1999), and Rad51 D
(Smiraldo et al, 2005) mutants was delayed on p53-null
background. The role of p53 in the embryonic lethality
of Xrcc2
/
mutants remains controversial (Orii et al,
2006; Adam et al, 2007).
The meiosis specific the Rad51 paralogue Dmc1 is essential
for meiotic recombination in S.cerevisiae (Bishop et al,
1992). In mice, Dmc1 expression is restricted to the testis
and ovaries, and null Dmc1 mutants are sterile and exhibit
arrested gametogenesis in the first meiotic prophase (Pittman
et al, 1998; Yoshida et al, 1998). These results demonstrate
the conservation of the essential meiotic function of Dmc1.
Rad54 is another member of the Rad52 epistasis group
involved in HR. Similar to rad51 and rad52 mutants, S.
cerevisiae deficient in rad54 are highly sensitive to IR. DT40
cells mutant for Rad54 are viable, hypersensitive to IR,
exhibit slow growth, decreased rate of Ig gene conversion,
and show a drastic decrease in gene-targeting efficiency
(Bezzubova et al, 1997). Rad54 paralogues exist and include
rdh54 in S.cerevisiae and Rad54B in mammalian cells. In
contrast to rad54 mutants in S.cerevisiae,rdh54 mutants are
not sensitive to IR and show no defects in mitosis (Shinohara
et al, 1997). However, meiotic recombination was affected in
rdh54 mutants and was completely defective in S.cerevisiae
lacking the two Rad54 orthologues (Shinohara et al, 1997).
Homozygous mutant mice lacking Rad54 and/or its para-
logue Rad54B are viable and ES cells deficient in Rad54 or
Rad54B are hypersensitive to IR, MMS, and MMC (Essers
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 597
et al, 1997; Bross et al, 2003; Wesoly et al, 2006). Rad54
/
,
but not Rad54B
/
, ES cells have a 3- to 40-fold decrease in
HR as assessed by gene targeting. Dual inactivation of both
paralogues further impairs HR, as double mutant ES cells
show a further 10-fold reduction in their gene targeting
efficiency compared with Rad54-null ES cells. Although
Rad54 is important for HR, mice deficient in Rad54 and/or
Rad54B do not have a predisposition to cancer.
BRCA1 and BRCA2, the early-onset breast cancer-
susceptibility genes, have also been demonstrated to partake
in HR-mediated DSB repair. Germline mutations of BRCA1 or
BRCA2 predispose women to familial human breast and
ovarian cancer (Narod and Foulkes, 2004). Individuals with
germline mutations for BRCA1 or BRCA2 also have increased
risk for other cancers, including prostate cancer. Although the
BRCA1 gene has important functions in DNA-damage signal-
ling/repair it is also involved in other cellular processes,
including transcription, ubiquitination, oestrogen receptor
signalling, and chromatin remodelling. In response to DSBs,
BRCA1 is phosphorylated by ATM, ATR, and Chk2, under-
scoring its functional regulation by other molecules involved
in DNA-damage checkpoints. BRCA2 functions in the loading
of the HR protein RAD51 during filament formation. It
directly binds to RAD51 and its phosphorylation on Ser3291
inhibits this binding (Esashi et al, 2005). In addition to its role
in HR, recent studies have suggested a role for BRCA2 in the
stabilization of stalled RFs (Lomonosov et al, 2003).
Homozygous mutations of murine Brca1 resulted in em-
bryonic lethality. (Gowen et al, 1996; Hakem et al, 1996; Liu
et al, 1996; Ludwig et al, 1997; Xu et al, 2001). Premature
death of Brca1-mutant embryos is likely due to accumulation
of damaged DNA and activation of DNA-damage checkpoints,
including Chk2–p53. Inactivation of p53 delayed the embryo-
nic lethality of Brca1-null mutants (Hakem et al, 1997;
Ludwig et al, 1997) and completely rescued survival of
Brca1 hypomorphic mutants (Xu et al, 2001). Studies of
hypomorphic and conditional null Brca1 mutant females
demonstrated that similar to humans, Brca1 mutations
increase the risk for mammary cancer (Xu et al, 1999;
McPherson et al, 2004b). Interestingly, inactivation of p53,
or Chk2, drastically facilitates development of mammary
tumors on Brca1-mutant backgrounds (Xu et al, 1999;
McPherson et al, 2004b). A targeted Brca1-null mutation to
the T-cell lineage resulted in increased genomic instability,
apoptosis, cell-cycle arrest, and a drastic depletion of the
T-cell lineage (Mak et al, 2000). Interestingly, development
of Brca1-null T-cells was completely rescued in p53-or
Chk2-mutant backgrounds, and this rescue was at the
expense of increased genomic instability and increased risk
for tumorigenesis (McPherson et al, 2004b).
BRCA1 plays a major role in the DNA-damage response as
it interacts with the MRN complex and g-H2AX at the site of
DSB. Likewise, it is also required for the recruitment of other
proteins such as Rad51, BRCA2, and BARD1 to sites of DSBs
(Greenberg et al, 2006). HR, as measured by the efficiency of
gene targeting in ES cells, was drastically reduced in Brca1
hypomorphic mutants (Moynahan et al, 1999). Other defects,
including compromised telomeres integrity and male sterility,
were also associated with Brca1 mutations (Xu et al, 2001;
McPherson et al, 2006).
Similar to Brca1 and Rad51 mutants, null mutations for
Brca2 result in early mouse embryonic lethality and impaired
HR (Ludwig et al, 1997; Suzuki et al, 1997; Moynahan et al,
2001). Furthermore, loss of Brca2 in murine cells resulted in
increased genomic instability and activation of p53. A small
subset of Brca2 hypomorphic mutants survived embryonic
development (Connor et al, 1997; Friedman et al, 1998).
These mice were growth defective, sterile, and predisposed
for thymic lymphoma, and their cells were sensitive to DNA
damaging agents, including IR, MMS, and UV radiation.
Conditional mutant strains have also been generated using
the Cre/loxP system and similar to Brca1 mutation, dual
inactivation of Brca2 and p53 in mammary epithelial cells
resulted in increased frequency and decreased latency for
mammary tumors (Jonkers et al, 2001).
Beside the proteins described above, efficient HR requires
other proteins such as members of the family of RecQ
helicases (Hickson, 2003). The human family of RecQ heli-
cases includes BLM, REQ, WRN, RECQL4, and RECQL5, and
is important for unwinding DNA, repairing stalled RFs,
promoting HR, and maintaining overall genomic integrity.
Mutations of some of these helicases are associated with rare
human syndromes (Hickson, 2003). Thus, the Werner syn-
drome (WS) and the Rothmund Thomson syndrome (RTS)
are associated with mutations of human WRN and RECQL4
genes, respectively (Figure 2). WS features include premature
ageing, short stature, and cancer predisposition. Clinical
features of RTS include short stature, skin pigmentation
changes, skin atrophy, and increased cancer predisposition.
In addition mutations of the human BLM gene are asso-
ciated with the Bloom syndrome (BS), a rare autosomal
recessive disorder characterized by growth defects, immune
deficiency, reduced fertility, and predisposition to a large
spectrum of cancers. Cells from BS patients have elevated
sister-chromatid exchange (SCE) and genomic instability.
Three mouse models for Blm mutation have been generated,
but only one was viable (Chester et al, 1998; Luo et al, 2000;
Goss et al, 2002). Similar to BS patients, viable Blm mice are
prone to a wide range of tumors, and cells from these mutant
mice show elevated levels of SCE, a hallmark cellular phe-
notype of BS. In addition, Blm deficiency leads to increased
mitotic recombination and somatic loss of heterozygosity
(Luo et al, 2000).
While most steps essential for HR in eukaryotes have been
well characterized, the identification of the resolvase(s)
required for the resolution of Holliday junctions (HJs) during
mitotic and meiotic recombination turned out to be very
challenging. Resolution of HJs is critical for HR and is
mediated in E.coli by the resolvase RuvC (West, 1997);
however, the search for true HJ resolvase(s) in eukaryotes
is still ongoing.
Yeast Mus81 (MMS, UV-sensitive, clone 81) with its partner
Eme1 (Schizosaccharomyces pombe) or MMS4 (S.cerevisiae)
forms a structure-specific endonuclease important for DNA-
damage repair (Osman and Whitby, 2007). Yeast mus81,
eme1,ormms4 mutants show increased sensitivity to DNA-
damaging agents that interfere with normal progression of
RFs caused by agents such as UV radiation, MMS, and
camptothecin, but their sensitivity to IR is not affected
(Osman and Whitby, 2007) . These mutants also show
meiotic defects supporting a role for this endonuclease
in this process.
In vitro studies indicate that yeast and mammalian Mus81,
with their partner Eme1/Mms4, efficiently cleaves various
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization598
DNA structures that mimic RFs, D-loops, and nicked HJs, but
the cleavage activity of intact HJs was weak (Osman and
Whitby, 2007). Thus, at least in mammalian cells, Mus81 with
Eme1 or with Eme2, another eme1 homologue, forms a
heterodimeric 30-flap/RF endonucleases that process stalled
RFs and recombination intermediates, but does not possess
typical characteristics of an HJ resolvase (Abraham et al,
2003; Ciccia et al, 2003, 2007; Ogrunc and Sancar, 2003).
Gene-targeted inactivation of Mus81 or Eme1 in mouse and
human cells increased sensitivity to ICL, hydroxyurea, but
not UV or IR and enhanced spontaneous and MMC-induced
genomic instability (Abraham et al, 2003; McPherson et al,
2004a). Mus81 was shown to be important for generating
ICL-induced DSBs, as well as for mediating the restart of
stalled or blocked RFs (Hanada et al, 2006). Our studies of a
mouse model for Mus81-null mutation demonstrated that
Mus81 is a haploinsufficient tumor suppressor (McPherson
et al, 2004a). Heterozygous and homozygous Mus81 mutants
show cancer predisposition, particularly to T- and B-cell
lymphomas. The MMC hypersensitivity of Mus81-mutant
cells and mice is p53-dependent (Pamidi et al, 2007).
Interestingly, on a p53-null background, Mus81-mutant
mice have more genomic instability and develop sarcomas
and multiple different tumors with short latency and high
penetrance, demonstrating a role for p53 in suppressing
cancer associated with Mus81 mutation (Pamidi et al,
2007). Both Mus81- and Mus81p53-mutant mice are fertile,
thus failing to support a requirement for Mus81 in meiotic
recombination. A second Mus81-mutant strain (Mus81
D912
)
showed increased genomic instability but failed to show
increased tumorigenesis (Dendouga et al, 2005). A role in
cancer for Eme1 or Eme2 requires further investigations.
Recent studies have implicated the mammalian RAD51
paralogues RAD51C–XRCC3 in HJ resolution (Liu et al,
2004). Rad51C–XRCC3 binds HJs in vitro and cell extracts
from Rad51C- or XRCC3-deficient hamster cells, exhibit low
HJ resolvase activities. In addition, depletion of RAD51C from
HeLa cell extracts strongly impairs in vitro HJ branch migra-
tion and resolution. Similar to other Rad51 paralogues, null
mutation of mouse Rad51C results in embryonic lethality.
However, a viable mouse model carrying intronic integration
of a Neomycine selection cassette resulting in decreased
Rad51 expression has been recently reported (Kuznetsov
et al, 2007). About 40% of mutant males and 10% of mutant
females were infertile. Cell extracts from MEFs null for
Rad51C and p53 show reduced in vitro HJ resolvase activities
compared with controls. Thus, the current data support a role
for the mammalian RAD51C in HJ resolution; however,
further studies are still required to establish RAD51C as
a typical HJ resolvase and to demonstrate whether other
mammalian HJ resolvases also exist.
Thus, the human syndromes and early onset of breast
cancer, and the developmental defects, increased genomic
instability and tumorigenesis associated with impaired HR in
mouse models, all demonstrate the in vivo requirement for
HR-mediated DNA-damage repair.
The NHEJ repair pathway
NHEJ is the predominant pathway of DSBs repair in mam-
malian cells (Figure 1; Kanaar et al, 2008). This repair path-
way is active especially at the G1, but is error prone. NHEJ is
also essential for T-cell receptor-a/band Ig V(D)J recombina-
tion, and thus this repair pathway is required for the devel-
opment of the T and B-cell repertoires. The core protein
components of the mammalian NHEJ include the Ku subunits
(Ku70 and Ku80), DNA–PKcs, XRCC4, DNA ligase IV (LigIV),
Artemis, and the recently identified Cernunnos–XLF (also
known as NHEJ1).
DNA–PK is composed of the catalytic subunit DNA–PKcs
and the heterodimer Ku70/Ku80, important for DNA end
binding. DNA–PKc is a serine/threonine kinase that is
activated following its recruitment by Ku70/Ku80 to sites of
DSBs. Active DNA–PKcs autophosphorylate themselves as
Table VI Examples of mouse models for the NHEJ repair pathway
Genotype Developmental defects Fertility
defects
Spontaneous
tumorigenesis
References
DNA-PKcs/T and B-cell development arrested
at early progenitor stages
None Not affected (Gao et al, 1998a; Taccioli
et al, 1998; Kurimasa et al,
1999)
Ku70/Growth retardation and early
arrest of T and B-cell development
None Thymic lymphomas (Gu et al, 1997; Ouyang
et al, 1997)
Ku80/Growth retardation and early
arrest of T and B-cell development
None Not affected (Nussenzweig et al, 1997)
Ku80/p53/Growth retardation and block of
T and B-cell development
None Early onset of pro
B-cell lymphomas
(Difilippantonio et al,
2000)
Artemis/Developmental arrest at early
T and B cell progenitor stages
None None (Rooney et al, 2002;
Li et al, 2005)
Artemis/p53/Developmental arrest at early
T and B cell progenitor stages
None Pro-B cell
lymphomas
(Rooney et al, 2004)
LigIV
Y288C
Growth retardation and progres-
sive loss of haematopoeitic stem
cells
NA None (Nijnik et al, 2007)
LigIV/Late embryonic lethality NA NA (Barnes et al, 1998;
Frank et al, 1998)
LigIV/p53/Viable but growth retarded None Early onset of pro
B-cell lymphomas
(Frank et al, 2000)
Xrcc4/Late embryonic lethality NA NA (Gao et al, 1998b)
Xrcc4/p53/Viable but growth retarded None Early onset of pro
B-cell lymphomas
(Gao et al, 2000)
NA, not applicable.
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 599
well as several other targets, including the Ku subunits, p53,
H2AX, Artemis, XRCC4, and WRN (Collis et al, 2005).
The important role of DNA–PKcs in vivo became evident
following the identification of its mutation in the severe
combined immunodeficient ‘SCID’ mice (Bosma et al, 1983;
Blunt et al, 1995; Kirchgessner et al, 1995). SCID mice exhibit
impaired V(D)J recombination and have arrested T- and
B-cell development at early progenitor stages. The role of
DNA–PKc mutation in SCID phenotypes was confirmed in
DNA–PKc-null mice obtained by gene targeting (Table VI; Gao
et al, 1998a; Taccioli et al, 1998; Kurimasa et al, 1999). These
mutants are severely immunodeficient, have impaired V(D)J
coding joining but normal signal joining, and their
T- and B-cell development are blocked at early progenitor
stages. Moreso, MEFs deficient for DNA–PKcs are hypersen-
sitive to IR, further supporting its requirement for DSB repair.
Similar to DNA–PKcs, targeted inactivation of Ku70 or
Ku80 in mice results in SCID phenotypes associated with
early arrest of T- and B-cell development, although the T-cell
arrest in Ku70 mice is leaky (Nussenzweig et al, 1996; Gu
et al, 1997; Ouyang et al, 1997). However, in contrast to
DNA–PKcs mice, null mutants for Ku70 or Ku80 showed
growth retardation and their cells were impaired in both
V(D)J coding and recombination signal (RS) end joining.
Inactivation of either Ku70 or Ku80 resulted in elevated IR
sensitivity but did not affect mice fertility. Taken together,
these data demonstrate the essential role of DNA–PK in NHEJ
and V(D)J recombination.
Increased apoptosis and proliferative arrest of pro-B cells
in Ku80
/
mice were rescued by a p53-null background
(Difilippantonio et al, 2000). Genomic instability associated
with Ku80 mutation was significantly increased in the ab-
sence of p53, and the double mutants developed pro B-cell
lymphomas with 100% penetrance and died within 3 months
of age (Nussenzweig et al, 1997; Difilippantonio et al, 2000).
Similarly, the incidence of thymic lymphomas was increased
in Ku70-null mice (Gu et al, 1997) and inactivation of p53 on
DNA–Pk
scid
-mutant background resulted in the rapid onset of
lymphomas/leukaemia (Guidos et al, 1996).
The nuclease Artemis is a phosphorylation target for
DNA–PK and forms a complex with DNA–PKcs. This complex
is important for the hairpin-opening step of V(D)J recombi-
nation and for the 50and 30overhang terminal end processing
in NHEJ. ARTEMIS mutation causes the severe combined
immunodeficiency with sensitivity to ionizing radiation
(RS-SCID), a rare human disorder (Figure 2; O’Driscoll and
Jeggo, 2006). Similar to DNA–PKc mutants, mice deficient for
Artemis are viable, have normal size, and suffer severe
combined immunodeficiency associated with developmental
arrest at early T- and B-cell progenitor stages (Rooney et al,
2002; Li et al, 2005). Artemis deficiency impaired coding but
not RS joining. Artemis-mutant MEFs are radiosensitive and
exhibit increased chromosomal instability. This spontaneous
loss of genomic integrity is likely the basis for increased
tumorigenesis associated with dual inactivation of Artemis
and p53 (Rooney et al, 2004; Woo et al, 2007).
Following the terminal end-processing step of NHEJ,
LigIV/XRCC4 complex serves to perform the ligation and
final step of NHEJ. Hypomorphic mutations of human LigIV
are associated with the hereditary autosomal LigIV syndrome
(Figure 2; O’Driscoll and Jeggo, 2006). This syndrome is very
rare, as only eight patients have been identified so far. This
syndrome is characterized by SCID phenotype, growth
defects, microcephaly, radiosensitivity, and leukaemia.
Recently, Cernunnos–XLF, a novel NHEJ factor that interacts
with the XRCC4–DNA LigIV complex, has been identified
(Ahnesorg et al, 2006; Buck et al, 2006). Mutation of
Cernunnos–XLF was found to be associated with a
rare inherited human syndrome characterized by growth
retardation, microcephaly, severe immunodeficiency, and
radiosensitivity (Figure 2).
Null mutation of LigIV or Xrcc4 in mouse models results in
late embryonic lethality associated with extensive apoptosis
in the embryonic central nervous system (Barnes et al, 1998;
Frank et al, 1998; Gao et al, 1998b). V(D)J joining does not
occur, lymphopoiesis is blocked, and MEFs deficient for LigIV
or Xrcc4 are radiosensitive, growth defective and enter sene-
scence prematurely. p53 inactivation rescued the extensive
apoptosis in the central nervous system, the defective pro-
liferation/senescence and the embryonic lethality associated
with LigIV or Xrcc4 mutation (Frank et al, 2000; Gao et al,
2000). However, p53 inactivation did not rescue defective
V(D)J recombination or lymphocyte development of LigIV-or
Xrcc4-null mutants. In addition, on a p53-null background,
both LigIV- and Xrcc4-null mutants were growth-retarded
and developed pro-B lymphomas with short latency. More
recently, the hypomorphic mutation LigIV
Y288C
has been
reported to lead to growth retardation, progressive loss of
haematopoeitic stem cells, and immunodeficiency (Nijnik
et al, 2007). In addition, mice carrying a targeted Xrcc4
mutation to neuronal progenitors demonstrated increased
genomic instability and predisposition for medulloblastomas
(Yan et al, 2006). Although mice deficient for Cernunnos–XLF
have not yet been reported, ES cells null for this gene are
radiosensitive, show defective V(D)J coding and RS joining,
and accumulate spontaneous genomic instability (Zha et al,
2007). These data suggest that similar to other NHEJ
core components, inactivation of Cernunnos–XLF could
potentially lead to cancer development.
The radiosensitivity, genomic instability, immuno-
deficiency, growth retardation, embryonic development, and
cancer predisposition associated with defective NHEJ all
demonstrate the major role this DNA-damage repair pathway
plays in vivo.
Concluding remarks
Although there exist numerous proteins employed by several
DNA-repair pathways in response to DNA damage, loss or
partial inactivation of only one of these proteins can have
extremely devastating consequences. Conversely, there are
some mutated DNA-damage repair proteins that go unnoticed
due to a lack of functional consequences at both the cellular
and organism level. This is likely due to genetic redundancy
and compensatory mechanisms of repair. The onset of var-
ious syndromes, increased cancer predisposition, immuno-
deficiency, and neurological defects associated with impaired
DNA-damage repair, are all strong indications for the require-
ment of a tight regulation of these pathways. Our current
knowledge of the mechanisms that regulate DNA repair
has grown significantly over the past years. In addition to
improving diagnosis, this overwhelming knowledge of
the mechanism of DNA-damage repair and DNA-damage
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization600
checkpoint will likely contribute to a better design for both
drugs and therapies for diseases, such as cancer.
Our understanding of the DNA-repair mechanisms in hu-
mans, and how defects in these processes lead to human
syndromes and pathologies, has greatly benefited from stu-
dies in other organisms, including mice and yeast. Further
studies in these organisms are required to better assess the
effect of the amino-acid substitutions and point mutations of
DNA-repair genes that associate with human syndromes
and diseases. Characterization of the post-translational
modifications that control the function and stability of DNA-
repair proteins, such as phosphorylation and ubiquitination,
is essential for future manipulation of the repair machinery to
aid in better therapeutic responses.
Acknowledgements
We thank the members of the Hakem laboratory for helpful discus-
sions, and Amanda Fenton, Anne Hakem, Jacinth Abraham, and
Renato Cardoso for reviewing the paper. Research in the Hakem
laboratory was supported by the Canadian Institutes of Health
Research, the National Cancer Institute of Canada, the Cancer
Research Society, and the Leukemia & Lymphoma Society Of
Canada. We apologize to those whose work was not cited directly
owing to space limitations.
References
Aas PA, Otterlei M, Falnes PO, Vagbo CB, Skorpen F, Akbari M,
Sundheim O, Bjoras M, Slupphaug G, Seeberg E, Krokan HE
(2003) Human and bacterial oxidative demethylases repair
alkylation damage in both RNA and DNA. Nature 421: 859–863
Abraham J, Lemmers B, Hande MP, Moynahan ME, Chahwan C,
Ciccia A, Essers J, Hanada K, Chahwan R, Khaw AK, McPherson
P, Shehabeldin A, Laister R, Arrowsmith C, Kanaar R, West SC,
Jasin M, Hakem R (2003) Eme1 is involved in DNA damage
processing and maintenance of genomic stability in mammalian
cells. EMBO J 22: 6137–6147
Adam J, Deans B, Thacker J (2007) A role for Xrcc2 in the early
stages of mouse development. DNA Repair (Amst) 6: 224–234
Ahnesorg P, Smith P, Jackson SP (2006) XLF interacts with the
XRCC4–DNA ligase IV complex to promote DNA nonhomologous
end-joining. Cell 124: 301–313
Andressoo JO, Mitchell JR, de Wit J, Hoogstraten D, Volker M,
Toussaint W, Speksnijder E, Beems RB, van Steeg H, Jans J, de
Zeeuw CI, Jaspers NG, Raams A, Lehmann AR, Vermeulen W,
Hoeijmakers JH, van der Horst GT (2006) An Xpd mouse model
for the combined xeroderma pigmentosum/Cockayne syndrome
exhibiting both cancer predisposition and segmental progeria.
Cancer Cell 10: 121–132
Baker SM, Bronner CE, Zhang L, Plug AW, Robatzek M, Warren G,
Elliott EA, Yu J, Ashley T, Arnheim N, Flavell RA, Liskay RM
(1995) Male mice defective in the DNA mismatch repair gene
PMS2 exhibit abnormal chromosome synapsis in meiosis. Cell 82:
309–319
Baker SM, Plug AW, Prolla TA, Bronner CE, Harris AC, Yao X,
Christie DM, Monell C, Arnheim N, Bradley A, Ashley T, Liskay
RM (1996) Involvement of mouse Mlh1 in DNA mismatch repair
and meiotic crossing over. Nat Genet 13: 336–342
Bardwell PD, Woo CJ, Wei K, Li Z, Martin A, Sack SZ, Parris T,
Edelmann W, Scharff MD (2004) Altered somatic hypermutation
and reduced class-switch recombination in exonuclease 1-mutant
mice. Nat Immunol 5: 224–229
Barnes DE, Stamp G, Rosewell I, Denzel A, Lindahl T (1998)
Targeted disruption of the gene encoding DNA ligase IV leads to
lethality in embryonic mice. Curr Biol 8: 1395–1398
Becker K, Gregel CM, Kaina B (1997) The DNA repair protein
O6-methylguanine–DNA methyltransferase protects against skin
tumor formation induced by antineoplastic chloroethylnitro-
sourea. Cancer Res 57: 3335–3338
Bender CF, Sikes ML, Sullivan R, Huye LE, Le Beau MM, Roth DB,
Mirzoeva OK, Oltz EM, Petrini JH (2002) Cancer predisposition
and hematopoietic failure in Rad50(S/S) mice. Genes Dev 16:
2237–2251
Bezzubova O, Silbergleit A, Yamaguchi-Iwai Y, Takeda S, Buerstedde
JM (1997) Reduced X-ray resistance and homologous recombina-
tion frequencies in a RAD54/mutant of the chicken DT40 cell
line. Cell 89: 185–193
Bishop DK, Park D, Xu L, Kleckner N (1992) DMC1: a meiosis-
specific yeast homolog of E.coli recA required for recombination,
synaptonemal complex formation, and cell cycle progression.
Cell 69: 439–456
Blunt T, Finnie NJ, Taccioli GE, Smith GC, Demengeot J, Gottlieb
TM, Mizuta R, Varghese AJ, Alt FW, Jeggo PA, Jackson SP (1995)
Defective DNA-dependent protein kinase activity is linked to
V(D)J recombination and DNA repair defects associated with
the murine scid mutation. Cell 80: 813–823
Bosma GC, Custer RP, Bosma MJ (1983) A severe combined
immunodeficiency mutation in the mouse. Nature 301: 527–530
Bross L, Wesoly J, Buerstedde JM, Kanaar R, Jacobs H (2003)
Somatic hypermutation does not require Rad54 and Rad54B-
mediated homologous recombination. Eur J Immunol 33:
352–357
Buck D, Malivert L, de Chasseval R, Barraud A, Fondaneche MC,
Sanal O, Plebani A, Stephan JL, Hufnagel M, le Deist F, Fischer A,
Durandy A, de Villartay JP, Revy P (2006) Cernunnos, a novel
nonhomologous end-joining factor, is mutated in human immu-
nodeficiency with microcephaly. Cell 124: 287–299
Cang Y, Zhang J, Nicholas SA, Bastien J, Li B, Zhou P, Goff SP
(2006) Deletion of DDB1 in mouse brain and lens leads to
p53-dependent elimination of proliferating cells. Cell 127:
929–940
Cheadle JP, Sampson JR (2007) MUTYH-associated polyposis—
from defect in base excision repair to clinical genetic testing.
DNA Repair (Amst) 6: 274–279
Chen PC, Dudley S, Hagen W, Dizon D, Paxton L, Reichow D, Yoon
SR, Yang K, Arnheim N, Liskay RM, Lipkin SM (2005)
Contributions by MutL homologues Mlh3 and Pms2 to DNA
mismatch repair and tumor suppression in the mouse. Cancer
Res 65: 8662–8670
Chester N, Kuo F, Kozak C, O’Hara CD, Leder P (1998) Stage-
specific apoptosis, developmental delay, and embryonic lethality
in mice homozygous for a targeted disruption in the murine
Bloom’s syndrome gene. Genes Dev 12: 3382–3393
Chipuk JE, Green DR (2006) Dissecting p53-dependent apoptosis.
Cell Death Differ 13: 994–1002
Ciccia A, Constantinou A, West SC (2003) Identification and
characterization of the human mus81–eme1 endonuclease.
J Biol Chem 278: 25172–25178
Ciccia A, Ling C, Coulthard R, Yan Z, Xue Y, Meetei AR, Laghmani el
H, Joenje H, McDonald N, de Winter JP, Wang W, West SC (2007)
Identification of FAAP24, a Fanconi anemia core complex protein
that interacts with FANCM. Mol Cell 25: 331–343
Collado M, Blasco MA, Serrano M (2007) Cellular senescence in
cancer and aging. Cell 130: 223–233
Collis SJ, DeWeese TL, Jeggo PA, Parker AR (2005) The life and
death of DNA–PK. Oncogene 24: 949–961
Connor F, Bertwistle D, Mee PJ, Ross GM, Swift S, Grigorieva E,
Tybulewicz VL, Ashworth A (1997) Tumorigenesis and a DNA
repair defect in mice with a truncating Brca2 mutation. Nat Genet
17: 423–430
de Boer J, de Wit J, van Steeg H, Berg RJ, Morreau H, Visser P,
Lehmann AR, Duran M, Hoeijmakers JH, Weeda G (1998a) A
mouse model for the basal transcription/DNA repair syndrome
trichothiodystrophy. Mol Cell 1: 981–990
de Boer J, Donker I, de Wit J, Hoeijmakers JH, Weeda G (1998b)
Disruption of the mouse xeroderma pigmentosum group D DNA
repair/basal transcription gene results in preimplantation lethal-
ity. Cancer Res 58: 89–94
de Boer J, van Steeg H, Berg RJ, Garssen J, de Wit J, van Oostrum
CT, Beems RB, van der Horst GT, van Kreijl CF, de Gruijl FR,
Bootsma D, Hoeijmakers JH, Weeda G (1999) Mouse model for
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 601
the DNA repair/basal transcription disorder trichothiodystrophy
reveals cancer predisposition. Cancer Res 59: 3489–3494
de Vries SS, Baart EB, Dekker M, Siezen A, de Rooij DG, de Boer P,
te Riele H (1999) Mouse MutS-like protein Msh5 is required for
proper chromosome synapsis in male and female meiosis. Genes
Dev 13: 523–531
de Wind N, Dekker M, Berns A, Radman M, te Riele H (1995)
Inactivation of the mouse Msh2 gene results in mismatch repair
deficiency, methylation tolerance, hyperrecombination, and
predisposition to cancer. Cell 82: 321–330
Deans B, Griffin CS, O’Regan P, Jasin M, Thacker J (2003)
Homologous recombination deficiency leads to profound genetic
instability in cells derived from Xrcc2-knockout mice. Cancer Res
63: 8181–8187
Dendouga N, Gao H, Moechars D, Janicot M, Vialard J, McGowan
CH (2005) Disruption of murine Mus81 increases genomic in-
stability and DNA damage sensitivity but does not promote
tumorigenesis. Mol Cell Biol 25: 7569–7579
Difilippantonio MJ, Zhu J, Chen HT, Meffre E, Nussenzweig MC,
Max EE, Ried T, Nussenzweig A (2000) DNA repair protein Ku80
suppresses chromosomal aberrations and malignant transforma-
tion. Nature 404: 510–514
Dumenco LL, Allay E, Norton K, Gerson SL (1993) The prevention
of thymic lymphomas in transgenic mice by human O6-alkylgua-
nine–DNA alkyltransferase. Science 259: 219–222
Dumon-Jones V, Frappart PO, Tong WM, Sajithlal G, Hulla W,
Schmid G, Herceg Z, Digweed M, Wang ZQ (2003) Nbn
heterozygosity renders mice susceptible to tumor formation and
ionizing radiation-induced tumorigenesis. Cancer Res 63:
7263–7269
Duncan T, Trewick SC, Koivisto P, Bates PA, Lindahl T, Sedgwick B
(2002) Reversal of DNA alkylation damage by two human
dioxygenases. Proc Natl Acad Sci USA 99: 16660–16665
Edelmann W, Cohen PE, Kane M, Lau K, Morrow B, Bennett S,
Umar A, Kunkel T, Cattoretti G, Chaganti R, Pollard JW,
Kolodner RD, Kucherlapati R (1996) Meiotic pachytene arrest in
MLH1-deficient mice. Cell 85: 1125–1134
Edelmann W, Cohen PE, Kneitz B, Winand N, Lia M, Heyer J,
Kolodner R, Pollard JW, Kucherlapati R (1999) Mammalian MutS
homologue 5 is required for chromosome pairing in meiosis. Nat
Genet 21: 123–127
Edelmann W, Umar A, Yang K, Heyer J, Kucherlapati M, Lia M,
Kneitz B, Avdievich E, Fan K, Wong E, Crouse G, Kunkel T, Lipkin
M, Kolodner RD, Kucherlapati R (2000) The DNA mismatch
repair genes Msh3 and Msh6 cooperate in intestinal tumor
suppression. Cancer Res 60: 803–807
Edelmann W, Yang K, Umar A, Heyer J, Lau K, Fan K, Liedtke W,
Cohen PE, Kane MF, Lipford JR, Yu N, Crouse GF, Pollard JW,
Kunkel T, Lipkin M, Kolodner R, Kucherlapati R (1997) Mutation
in the mismatch repair gene Msh6 causes cancer susceptibility.
Cell 91: 467–477
Engelward BP, Weeda G, Wyatt MD, Broekhof JL, de Wit J, Donker
I, Allan JM, Gold B, Hoeijmakers JH, Samson LD (1997) Base
excision repair deficient mice lacking the Aag alkyladenine DNA
glycosylase. Proc Natl Acad Sci USA 94: 13087–13092
Esashi F, Christ N, Gannon J, Liu Y, Hunt T, Jasin M, West SC (2005)
CDK-dependent phosphorylation of BRCA2 as a regulatory me-
chanism for recombinational repair. Nature 434: 598–604
Essers J, Hendriks RW, Swagemakers SM, Troelstra C, de Wit J,
Bootsma D, Hoeijmakers JH, Kanaar R (1997) Disruption of
mouse RAD54 reduces ionizing radiation resistance and homo-
logous recombination. Cell 89: 195–204
Falnes PO, Bjoras M, Aas PA, Sundheim O, Seeberg E (2004)
Substrate specificities of bacterial and human AlkB proteins.
Nucleic Acids Res 32: 3456–3461
Felton KE, Gilchrist DM, Andrew SE (2007) Constitutive deficiency
in DNA mismatch repair. Clin Genet 71: 483–498
Frank KM, Sekiguchi JM, Seidl KJ, Swat W, Rathbun GA, Cheng HL,
Davidson L, Kangaloo L, Alt FW (1998) Late embryonic lethality
and impaired V(D)J recombination in mice lacking DNA ligase IV.
Nature 396: 173–177
Frank KM, Sharpless NE, Gao Y, Sekiguchi JM, Ferguson DO, Zhu C,
Manis JP, Horner J, DePinho RA, Alt FW (2000) DNA ligase IV
deficiency in mice leads to defective neurogenesis and embryonic
lethality via the p53 pathway. Mol Cell 5: 993–1002
Frappart PO, Tong WM, Demuth I, Radovanovic I, Herceg Z, Aguzzi
A, Digweed M, Wang ZQ (2005) An essential function for NBS1 in
the prevention of ataxia and cerebellar defects. Nat Med 11 :
538–544
Friedman LS, Thistlethwaite FC, Patel KJ, Yu VP, Lee H,
Venkitaraman AR, Abel KJ, Carlton MB, Hunter SM, Colledge
WH, Evans MJ, Ponder BA (1998) Thymic lymphomas in mice
with a truncating mutation in Brca2. Cancer Res 58: 1338–1343
Gao Y, Chaudhuri J, Zhu C, Davidson L, Weaver DT, Alt FW (1998a)
A targeted DNA–PKcs-null mutation reveals DNA-PK-indepen-
dent functions for KU in V(D)J recombination. Immunity 9:
367–376
Gao Y, Ferguson DO, Xie W, Manis JP, Sekiguchi J, Frank KM,
Chaudhuri J, Horner J, DePinho RA, Alt FW (2000) Interplay of
p53 and DNA-repair protein XRCC4 in tumorigenesis, genomic
stability and development. Nature 404: 897–900
Gao Y, Sun Y, Frank KM, Dikkes P, Fujiwara Y, Seidl KJ, Sekiguchi
JM, Rathbun GA, Swat W, Wang J, Bronson RT, Malynn BA,
Bryans M, Zhu C, Chaudhuri J, Davidson L, Ferrini R, Stamato T,
Orkin SH, Greenberg ME et al (1998b) A critical role for DNA end-
joining proteins in both lymphogenesis and neurogenesis. Cell 95:
891–902
Glassner BJ, Weeda G, Allan JM, Broekhof JL, Carls NH, Donker I,
Engelward BP, Hampson RJ, Hersmus R, Hickman MJ, Roth RB,
Warren HB, Wu MM, Hoeijmakers JH, Samson LD (1999) DNA
repair methyltransferase (Mgmt) knockout mice are sensitive to
the lethal effects of chemotherapeutic alkylating agents.
Mutagenesis 14: 339–347
Goss KH, Risinger MA, Kordich JJ, Sanz MM, Straughen JE, Slovek
LE, Capobianco AJ, German J, Boivin GP, Groden J (2002)
Enhanced tumor formation in mice heterozygous for Blm muta-
tion. Science 297: 2051–2053
Gowen LC, Johnson BL, Latour AM, Sulik KK, Koller BH (1996)
Brca1 deficiency results in early embryonic lethality characterized
by neuroepithelial abnormalities. Nat Genet 12: 191–194
Greenberg RA, Sobhian B, Pathania S, Cantor SB, Nakatani Y,
Livingston DM (2006) Multifactorial contributions to an acute
DNA damage response by BRCA1/BARD1-containing complexes.
Genes Dev 20: 34–46
Gu H, Marth JD, Orban PC, Mossmann H, Rajewsky K (1994)
Deletion of a DNA polymerase beta gene segment in T cells
using cell type-specific gene targeting. Science 265: 103–106
Gu Y, Seidl KJ, Rathbun GA, Zhu C, Manis JP, van der Stoep N,
Davidson L, Cheng HL, Sekiguchi JM, Frank K, Stanhope-Baker P,
Schlissel MS, Roth DB, Alt FW (1997) Growth retardation and
leaky SCID phenotype of Ku70-deficient mice. Immunity 7:
653–665
Guidos CJ, Williams CJ, Grandal I, Knowles G, Huang MT, Danska
JS (1996) V(D)J recombination activates a p53-dependent DNA
damage checkpoint in scid lymphocyte precursors. Genes Dev 10:
2038–2054
Hakem R, de la Pompa JL, Elia A, Potter J, Mak TW (1997) Partial
rescue of Brca1 (5–6) early embryonic lethality by p53 or p21 null
mutation. Nat Genet 16: 298–302
Hakem R, de la Pompa JL, Sirard C, Mo R, Woo M, Hakem A,
Wakeham A, Potter J, Reitmair A, Billia F, Firpo E, Hui CC,
Roberts J, Rossant J, Mak TW (1996) The tumor suppressor
gene Brca1 is required for embryonic cellular proliferation in
the mouse. Cell 85: 1009–1023
Hanada K, Budzowska M, Modesti M, Maas A, Wyman C, Essers J,
Kanaar R (2006) The structure-specific endonuclease Mus81–
Eme1 promotes conversion of interstrand DNA crosslinks into
double-strands breaks. EMBO J 25: 4921–4932
Harada YN, Shiomi N, Koike M, Ikawa M, Okabe M, Hirota S,
Kitamura Y, Kitagawa M, Matsunaga T, Nikaido O, Shiomi T
(1999) Postnatal growth failure, short life span, and early onset of
cellular senescence and subsequent immortalization in mice
lacking the xeroderma pigmentosum group G gene. Mol Cell
Biol 19: 2366–2372
Hickson ID (2003) RecQ helicases: caretakers of the genome. Nat
Rev Cancer 3: 169–178
Itoh T, Cado D, Kamide R, Linn S (2004) DDB2 gene disruption
leads to skin tumors and resistance to apoptosis after exposure to
ultraviolet light but not a chemical carcinogen. Proc Natl Acad Sci
USA 101: 2052–2057
Iwakuma T, Sakumi K, Nakatsuru Y, Kawate H, Igarashi H, Shiraishi
A, Tsuzuki T, Ishikawa T, Sekiguchi M (1997) High incidence of
nitrosamine-induced tumorigenesis in mice lacking DNA repair
methyltransferase. Carcinogenesis 18: 1631–1635
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization602
Jiricny J (2006) The multifaceted mismatch-repair system. Nat Rev
Mol Cell Biol 7: 335–346
Jonkers J, Meuwissen R, van der Gulden H, Peterse H, van der
Valk M, Berns A (2001) Synergistic tumor suppressor activity of
BRCA2 and p53 in a conditional mouse model for breast cancer.
Nat Genet 29: 418–425
Kanaar R, Wyman C, Rothstein R (2008) Quality control of
DNA break metabolism: in the ‘end’, it’s a good thing.
EMBO J 27: 581–588
Kane MF, Loda M, Gaida GM, Lipman J, Mishra R, Goldman H,
Jessup JM, Kolodner R (1997) Methylation of the hMLH1 pro-
moter correlates with lack of expression of hMLH1 in sporadic
colon tumors and mismatch repair-defective human tumor cell
lines. Cancer Res 57: 808–811
Kang J, Bronson RT, Xu Y (2002) Targeted disruption of NBS1
reveals its roles in mouse development and DNA repair. EMBO
J21: 1447–1455
Kirchgessner CU, Patil CK, Evans JW, Cuomo CA, Fried LM, Carter
T, Oettinger MA, Brown JM (1995) DNA-dependent kinase (p350)
as a candidate gene for the murine SCID defect. Science 267:
1178–1183
Klungland A, Rosewell I, Hollenbach S, Larsen E, Daly G, Epe B,
Seeberg E, Lindahl T, Barnes DE (1999) Accumulation of pre-
mutagenic DNA lesions in mice defective in removal of oxidative
base damage. Proc Natl Acad Sci USA 96: 13300–13305
Kneitz B, Cohen PE, Avdievich E, Zhu L, Kane MF, Hou Jr H,
Kolodner RD, Kucherlapati R, Pollard JW, Edelmann W (2000)
MutS homolog 4 localization to meiotic chromosomes is required
for chromosome pairing during meiosis in male and female mice.
Genes Dev 14: 1085–1097
Krogh BO, Symington LS (2004) Recombination proteins in yeast.
Annu Rev Genet 38: 233–271
Kucherlapati M, Yang K, Kuraguchi M, Zhao J, Lia M, Heyer J,
Kane MF, Fan K, Russell R, Brown AM, Kneitz B, Edelmann W,
Kolodner RD, Lipkin M, Kucherlapati R (2002) Haploinsufficiency
of Flap endonuclease (Fen1) leads to rapid tumor progression.
Proc Natl Acad Sci USA 99: 9924–9929
Kurimasa A, Ouyang H, Dong LJ, Wang S, Li X, Cordon-Cardo C,
Chen DJ, Li GC (1999) Catalytic subunit of DNA-dependent
protein kinase: impact on lymphocyte development and tumor-
igenesis. Proc Natl Acad Sci USA 96: 1403–1408
Kuznetsov S, Pellegrini M, Shuda K, Fernandez-Capetillo O, Liu Y,
Martin BK, Burkett S, Southon E, Pati D, Tessarollo L, West SC,
Donovan PJ, Nussenzweig A, Sharan SK (2007) RAD51C defi-
ciency in mice results in early prophase I arrest in males and
sister chromatid separation at metaphase II in females. J Cell Biol
176: 581–592
Leibeling D, Laspe P, Emmert S (2006) Nucleotide excision repair
and cancer. J Mol Histol 37: 225–238
Li L, Salido E, Zhou Y, Bhattacharyya S, Yannone SM, Dunn E,
Meneses J, Feeney AJ, Cowan MJ (2005) Targeted disruption of
the Artemis murine counterpart results in SCID and defective
V(D)J recombination that is partially corrected with bone marrow
transplantation. J Immunol 174: 2420–2428
Li Z, Scherer SJ, Ronai D, Iglesias-Ussel MD, Peled JU, Bardwell PD,
Zhuang M, Lee K, Martin A, Edelmann W, Scharff MD (2004)
Examination of Msh6- and Msh3-deficient mice in class switching
reveals overlapping and distinct roles of MutS homologues in
antibody diversification. J Exp Med 200: 47–59
Lim DS, Hasty P (1996) A mutation in mouse rad51 results in an
early embryonic lethal that is suppressed by a mutation in p53.
Mol Cell Biol 16: 7133–7143
Lipkin SM, Moens PB, Wang V, Lenzi M, Shanmugarajah D,
Gilgeous A, Thomas J, Cheng J, Touchman JW, Green ED,
Schwartzberg P, Collins FS, Cohen PE (2002) Meiotic arrest and
aneuploidy in MLH3-deficient mice. Nat Genet 31: 385–390
Liu CY, Flesken-Nikitin A, Li S, Zeng Y, Lee WH (1996) Inactivation
of the mouse Brca1 gene leads to failure in the morphogenesis of
the egg cylinder in early postimplantation development. Genes
Dev 10: 1835–1843
Liu Y, Masson JY, Shah R, O’Regan P, West SC (2004) RAD51C is
required for Holliday junction processing in mammalian cells.
Science 303: 243–246
Lomonosov M, Anand S, Sangrithi M, Davies R, Venkitaraman AR
(2003) Stabilization of stalled DNA replication forks by the
BRCA2 breast cancer susceptibility protein. Genes Dev 17:
3017–3022
Ludwig DL, MacInnes MA, Takiguchi Y, Purtymun PE, Henrie M,
Flannery M, Meneses J, Pedersen RA, Chen DJ (1998) A murine
AP-endonuclease gene-targeted deficiency with post-implantation
embryonic progression and ionizing radiation sensitivity. Mutat
Res 409: 17–29
Ludwig T, Chapman DL, Papaioannou VE, Efstratiadis A (1997)
Targeted mutations of breast cancer susceptibility gene homologs
in mice: lethal phenotypes of Brca1, Brca2, Brca1/Brca2,
Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev 11:
1226–1241
Luo G, Santoro IM, McDaniel LD, Nishijima I, Mills M, Youssoufian
H, Vogel H, Schultz RA, Bradley A (2000) Cancer predisposition
caused by elevated mitotic recombination in Bloom mice. Nat
Genet 26: 424–429
Luo G, Yao MS, Bender CF, Mills M, Bladl AR, Bradley A, Petrini JH
(1999) Disruption of mRad50 causes embryonic stem cell leth-
ality, abnormal embryonic development, and sensitivity to ioniz-
ing radiation. Proc Natl Acad Sci USA 96: 7376–7381
Maizels N (2005) Immunoglobulin gene diversification. Annu Rev
Genet 39: 23–46
Mak TW, Hakem A, McPherson JP, Shehabeldin A, Zablocki E,
Migon E, Duncan GS, Bouchard D, Wakeham A, Cheung A,
Karaskova J, Sarosi I, Squire J, Marth J, Hakem R (2000) Brcal
required for T cell lineage development but not TCR loci rearran-
gement. Nat Immunol 1: 77–82
Marsischky GT, Filosi N, Kane MF, Kolodner R (1996) Redundancy
of Saccharomyces cerevisiae MSH3 and MSH6 in MSH2-dependent
mismatch repair. Genes Dev 10: 407–420
McPherson JP, Hande MP, Poonepalli A, Lemmers B, Zablocki E,
Migon E, Shehabeldin A, Porras A, Karaskova J, Vukovic B,
Squire J, Hakem R (2006) A role for Brca1 in chromosome end
maintenance. Hum Mol Genet 15: 831–838
McPherson JP, Lemmers B, Chahwan R, Pamidi A, Migon E,
Matysiak-Zablocki E, Moynahan ME, Essers J, Hanada K,
Poonepalli A, Sanchez-Sweatman O, Khokha R, Kanaar R, Jasin
M, Hande MP, Hakem R (2004a) Involvement of mammalian
Mus81 in genome integrity and tumor suppression. Science 304:
1822–1826
McPherson JP, Lemmers B, Hirao A, Hakem A, Abraham J, Migon E,
Matysiak-Zablocki E, Tamblyn L, Sanchez-Sweatman O,
Khokha R, Squire J, Hande MP, Mak TW, Hakem R (2004b)
Collaboration of Brca1 and Chk2 in tumorigenesis. Genes Dev
18: 1144–1153
McWhir J, Selfridge J, Harrison DJ, Squires S, Melton DW (1993)
Mice with DNA repair gene (ERCC-1) deficiency have elevated
levels of p53, liver nuclear abnormalities and die before weaning.
Nat Genet 5: 217–224
Morales M, Theunissen JW, Kim CF, Kitagawa R, Kastan MB, Petrini
JH (2005) The Rad50S allele promotes ATM-dependent DNA
damage responses and suppresses ATM deficiency: implications
for the Mre11 complex as a DNA damage sensor. Genes Dev 19:
3043–3054
Moynahan ME, Chiu JW, Koller BH, Jasin M (1999) Brca1 controls
homology-directed DNA repair. Mol Cell 4: 511–518
Moynahan ME, Pierce AJ, Jasin M (2001) BRCA2 is required for
homology-directed repair of chromosomal breaks. Mol Cell 7:
263–272
Nakane H, Takeuchi S, Yuba S, Saijo M, Nakatsu Y, Murai H,
Nakatsuru Y, Ishikawa T, Hirota S, Kitamura Y, Kato Y, Tsunoda Y,
Miyauchi H, Horio T, Tokunaga T, Matsunaga T, Nikaido O,
Nishimune Y, Okada S, Okada S, Tanaka H (1995) High incidence
of ultraviolet-B-or chemical-carcinogen-induced skin tumours in
mice lacking the xeroderma pigmentosum group A gene. Nature
377: 165–168
Nakatsuru Y, Matsukuma S, Nemoto N, Sugano H, Sekiguchi M,
Ishikawa T (1993) O6-methylguanine–DNA methyltransferase
protects against nitrosamine-induced hepatocarcinogenesis. Proc
Natl Acad Sci USA 90: 6468–6472
Narod SA, Foulkes WD (2004) BRCA1 and BRCA2: 1994 and
beyond. Nat Rev Cancer 4: 665–676
Ng JM, Vermeulen W, van der Horst GT, Bergink S, Sugasawa K,
Vrieling H, Hoeijmakers JH (2003) A novel regulation mechanism
of DNA repair by damage-induced and RAD23-dependent stabi-
lization of xeroderma pigmentosum group C protein. Genes Dev
17: 1630–1645
Ng JM, Vrieling H, Sugasawa K, Ooms MP, Grootegoed JA, Vreeburg
JT, Visser P, Beems RB, Gorgels TG, Hanaoka F, Hoeijmakers JH,
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 603
van der Horst GT (2002) Developmental defects and male sterility
in mice lacking the ubiquitin-like DNA repair gene mHR23B. Mol
Cell Biol 22: 1233–1245
Niida H, Nakanishi M (2006) DNA damage checkpoints in mam-
mals. Mutagenesis 21: 3–9
Nijnik A, Woodbine L, Marchetti C, Dawson S, Lambe T, Liu C,
Rodrigues NP, Crockford TL, Cabuy E, Vindigni A, Enver T, Bell
JI, Slijepcevic P, Goodnow CC, Jeggo PA, Cornall RJ (2007) DNA
repair is limiting for haematopoietic stem cells during ageing.
Nature 447: 686–690
Nilsen H, Stamp G, Andersen S, Hrivnak G, Krokan HE, Lindahl T,
Barnes DE (2003) Gene-targeted mice lacking the Ung
uracil–DNA glycosylase develop B-cell lymphomas. Oncogene
22: 5381–5386
Nussenzweig A, Chen C, da Costa Soares V, Sanchez M, Sokol K,
Nussenzweig MC, Li GC (1996) Requirement for Ku80 in growth
and immunoglobulin V(D)J recombination. Nature 382: 551–555
Nussenzweig A, Sokol K, Burgman P, Li L, Li GC (1997)
Hypersensitivity of Ku80-deficient cell lines and mice to DNA
damage: the effects of ionizing radiation on growth, survival, and
development. Proc Natl Acad Sci USA 94: 13588–13593
O’Driscoll M, Jeggo PA (2006) The role of double-strand break
repair—insights from human genetics. Nat Rev Genet 7: 45–54
Ocampo MT, Chaung W, Marenstein DR, Chan MK, Altamirano A,
Basu AK, Boorstein RJ, Cunningham RP, Teebor GW (2002)
Targeted deletion of mNth1 reveals a novel DNA repair enzyme
activity. Mol Cell Biol 22: 6111–6121
Ogrunc M, Sancar A (2003) Identification and characterization of
human MUS81-MMS4 structure-specific endonuclease. J Biol
Chem 278: 21715–21720
Orii KE, Lee Y, Kondo N, McKinnon PJ (2006) Selective utilization of
nonhomologous end-joining and homologous recombination
DNA repair pathways during nervous system development. Proc
Natl Acad Sci USA 103: 10017–10022
Osman F, Whitby MC (2007) Exploring the roles of Mus81–
Eme1/Mms4 at perturbed replication forks. DNA Repair (Amst)
6: 1004–1017
Ouyang H, Nussenzweig A, Kurimasa A, Soares VC, Li X, Cordon-
Cardo C, Li W, Cheong N, Nussenzweig M, Iliakis G, Chen DJ,
Li GC (1997) Ku70 is required for DNA repair but not for T
cell antigen receptor gene recombination in vivo.J Exp Med 186:
921–929
Pamidi A, Cardoso R, Hakem A, Matysiak-Zablocki E, Poonepalli A,
Tamblyn L, Perez-Ordonez B, Hande MP, Sanchez O, Hakem R
(2007) Functional interplay of p53 and Mus81 in DNA damage
responses and cancer. Cancer Res 67: 8527–8535
Peltomaki P, Vasen HF (1997) Mutations predisposing to hereditary
nonpolyposis colorectal cancer: database and results of a colla-
borative study. The International Collaborative Group on
Hereditary Nonpolyposis Colorectal Cancer. Gastroenterology
113: 1146–1158
Petrini JH, Xiao Y, Weaver DT (1995) DNA ligase I mediates
essential functions in mammalian cells. Mol Cell Biol 15:
4303–4308
Pittman DL, Cobb J, Schimenti KJ, Wilson LA, Cooper DM, Brignull
E, Handel MA, Schimenti JC (1998) Meiotic prophase arrest with
failure of chromosome synapsis in mice deficient for Dmc1, a
germline-specific RecA homolog. Mol Cell 1: 697–705
Pittman DL, Schimenti JC (2000) Midgestation lethality in mice
deficient for the RecA-related gene, Rad51d/Rad51l3. Genesis 26:
167–173
Prolla TA, Baker SM, Harris AC, Tsao JL, Yao X, Bronner CE, Zheng
B, Gordon M, Reneker J, Arnheim N, Shibata D, Bradley A, Liskay
RM (1998) Tumour susceptibility and spontaneous mutation in
mice deficient in Mlh1, Pms1 and Pms2 DNA mismatch repair.
Nat Genet 18: 276–279
Puebla-Osorio N, Lacey DB, Alt FW, Zhu C (2006) Early embryonic
lethality due to targeted inactivation of DNA ligase III. Mol Cell
Biol 26: 3935–3941
Rada C, Ehrenstein MR, Neuberger MS, Milstein C (1998) Hot spot
focusing of somatic hypermutation in MSH2-deficient mice sug-
gests two stages of mutational targeting. Immunity 9: 135–141
Reitmair AH, Schmits R, Ewel A, Bapat B, Redston M, Mitri A,
Waterhouse P, Mittrucker HW, Wakeham A, Liu B, Thomason A,
Griesser H, Gallinger S, Ballhausen WG, Fishel R, Mak TW (1995)
MSH2 deficient mice are viable and susceptible to lymphoid
tumours. Nat Genet 11 : 64–70
Rijkers T, Van Den Ouweland J, Morolli B, Rolink AG, Baarends
WM, Van Sloun PP, Lohman PH, Pastink A (1998) Targeted
inactivation of mouse RAD52 reduces homologous recombina-
tion but not resistance to ionizing radiation. Mol Cell Biol 18:
6423–6429
Ringvoll J, Nordstrand LM, Vagbo CB, Talstad V, Reite K, Aas PA,
Lauritzen KH, Liabakk NB, Bjork A, Doughty RW, Falnes PO,
Krokan HE, Klungland A (2006) Repair deficient mice reveal
mABH2 as the primary oxidative demethylase for repairing
1meA and 3meC lesions in DNA. EMBO J 25: 2189–2198
Rooney S, Sekiguchi J, Whitlow S, Eckersdorff M, Manis JP, Lee C,
Ferguson DO, Alt FW (2004) Artemis and p53 cooperate to
suppress oncogenic N-myc amplification in progenitor B cells.
Proc Natl Acad Sci USA 101: 2410–2415
Rooney S, Sekiguchi J, Zhu C, Cheng HL, Manis J, Whitlow S,
DeVido J, Foy D, Chaudhuri J, Lombard D, Alt FW (2002) Leaky
Scid phenotype associated with defective V(D)J coding end
processing in Artemis-deficient mice. Mol Cell 10: 1379–1390
Sakamoto K, Tominaga Y, Yamauchi K, Nakatsu Y, Sakumi K,
Yoshiyama K, Egashira A, Kura S, Yao T, Tsuneyoshi M, Maki
H, Nakabeppu Y, Tsuzuki T (2007) MUTYH-null mice are suscep-
tible to spontaneous and oxidative stress induced intestinal
tumorigenesis. Cancer Res 67: 6599–6604
Sands AT, Abuin A, Sanchez A, Conti CJ, Bradley A (1995) High
susceptibility to ultraviolet-induced carcinogenesis in mice lack-
ing XPC. Nature 377: 162–165
Schmutte C, Sadoff MM, Shim KS, Acharya S, Fishel R (2001) The
interaction of DNA mismatch repair proteins with human exonu-
clease I. J Biol Chem 276: 33011–33018
Sedgwick B, Bates PA, Paik J, Jacobs SC, Lindahl T (2007) Repair of
alkylated DNA: recent advances. DNA Repair (Amst) 6: 429–442
Shinohara A, Ogawa H, Ogawa T (1992) Rad51 protein involved in
repair and recombination in S.cerevisiae is a RecA-like protein.
Cell 69: 457–470
Shinohara M, Shita-Yamaguchi E, Buerstedde JM, Shinagawa H,
Ogawa H, Shinohara A (1997) Characterization of the roles of the
Saccharomyces cerevisiae RAD54 gene and a homologue of
RAD54, RDH54/TID1, in mitosis and meiosis. Genetics 147:
1545–1556
Shu Z, Smith S, Wang L, Rice MC, Kmiec EB (1999) Disruption of
muREC2/RAD51L1 in mice results in early embryonic lethality
which can be partially rescued in a p53(/) background. Mol
Cell Biol 19: 8686–8693
Smiraldo PG, Gruver AM, Osborn JC, Pittman DL (2005) Extensive
chromosomal instability in Rad51d-deficient mouse cells. Cancer
Res 65: 2089–2096
Sonoda E, Sasaki MS, Buerstedde JM, Bezzubova O, Shinohara A,
Ogawa H, Takata M, Yamaguchi-Iwai Y, Takeda S (1998) Rad51-
deficient vertebrate cells accumulate chromosomal breaks prior
to cell death. EMBO J 17: 598–608
Su TT (2006) Cellular responses to DNA damage: one signal,
multiple choices. Annu Rev Genet 40: 187–208
Subba Rao K (2007) Mechanisms of disease: DNA repair defects and
neurological disease. Nat Clin Pract Neurol 3: 162–172
Sugo N, Aratani Y, Nagashima Y, Kubota Y, Koyama H (2000)
Neonatal lethality with abnormal neurogenesis in mice deficient
in DNA polymerase beta. EMBO J 19: 1397–1404
Sung P, Klein H (2006) Mechanism of homologous recombination:
mediators and helicases take on regulatory functions. Nat Rev Mol
Cell Biol 7: 739–750
Suzuki A, de la Pompa JL, Hakem R, Elia A, Yoshida R, Mo R,
Nishina H, Chuang T, Wakeham A, Itie A, Koo W, Billia P, Ho A,
Fukumoto M, Hui CC, Mak TW (1997) Brca2 is required for
embryonic cellular proliferation in the mouse. Genes Dev 11:
1242–1252
Symington LS (2002) Role of RAD52 epistasis group genes in
homologous recombination and double-strand break repair.
Microbiol Mol Biol Rev 66: 630–670, table of contents
Taccioli GE, Amatucci AG, Beamish HJ, Gell D, Xiang XH, Torres
Arzayus MI, Priestley A, Jackson SP, Marshak Rothstein A, Jeggo
PA, Herrera VL (1998) Targeted disruption of the catalytic subunit
of the DNA-PK gene in mice confers severe combined immuno-
deficiency and radiosensitivity. Immunity 9: 355–366
Takao M, Kanno S, Shiromoto T, Hasegawa R, Ide H, Ikeda S, Sarker
AH, Seki S, Xing JZ, Le XC, Weinfeld M, Kobayashi K, Miyazaki J,
Muijtjens M, Hoeijmakers JH, van der Horst G, Yasui A (2002)
Novel nuclear and mitochondrial glycosylases revealed by
DNA-damage repair
R Hakem
The EMBO Journal VOL 27 |NO 4 |2008 &2008 European Molecular Biology Organization604
disruption of the mouse Nth1 gene encoding an endonuclease III
homolog for repair of thymine glycols. EMBO J 21 : 3486–3493
Takata M, Sasaki MS, Sonoda E, Fukushima T, Morrison C, Albala
JS, Swagemakers SM, Kanaar R, Thompson LH, Takeda S (2000)
The Rad51 paralog Rad51B promotes homologous recombina-
tional repair. Mol Cell Biol 20: 6476–6482
Takata M, Sasaki MS, Tachiiri S, Fukushima T, Sonoda E, Schild D,
Thompson LH, Takeda S (2001) Chromosome instability and
defective recombinational repair in knockout mutants of the
five Rad51 paralogs. Mol Cell Biol 21: 2858–2866
Tauchi H, Kobayashi J, Morishima K, van Gent DC, Shiraishi T,
Verkaik NS, vanHeems D, Ito E, Nakamura A, Sonoda E, Takata
M, Takeda S, Matsuura S, Komatsu K (2002) Nbs1 is essential for
DNA repair by homologous recombination in higher vertebrate
cells. Nature 420: 93–98
Taylor AM, Groom A, Byrd PJ (2004) Ataxia–telangiectasia-like
disorder (ATLD)-its clinical presentation and molecular basis.
DNA Repair (Amst) 3: 1219–1225
Tebbs RS, Flannery ML, Meneses JJ, Hartmann A, Tucker JD,
Thompson LH, Cleaver JE, Pedersen RA (1999) Requirement for
the Xrcc1 DNA base excision repair gene during early mouse
development. Dev Biol 208: 513–529
Theunissen JW, Kaplan MI, Hunt PA, Williams BR, Ferguson DO, Alt
FW, Petrini JH (2003) Checkpoint failure and chromosomal
instability without lymphomagenesis in Mre11(ATLD1/ATLD1)
mice. Mol Cell 12: 1511–1523
Thompson LH, Schild D (2002) Recombinational DNA repair and
human disease. Mutat Res 509: 49–78
Thoms KM, Kuschal C, Emmert S (2007) Lessons learned from DNA
repair defective syndromes. Exp Dermatol 16: 532–544
Tian M, Jones DA, Smith M, Shinkura R, Alt FW (2004a) Deficiency
in the nuclease activity of xeroderma pigmentosum G in mice
leads to hypersensitivity to UV irradiation. Mol Cell Biol 24:
2237–2242
Tian M, Shinkura R, Shinkura N, Alt FW (2004b) Growth
retardation, early death, and DNA repair defects in mice deficient
for the nucleotide excision repair enzyme XPF. Mol Cell Biol 24:
1200–1205
Tishkoff DX, Boerger AL, Bertrand P, Filosi N, Gaida GM, Kane MF,
Kolodner RD (1997) Identification and characterization of
Saccharomyces cerevisiae EXO1, a gene encoding an exonuclease
that interacts with MSH2. Proc Natl Acad Sci USA 94: 7487–7492
Treuner K, Helton R, Barlow C (2004) Loss of Rad52 partially
rescues tumorigenesis and T-cell maturation in Atm-deficient
mice. Oncogene 23: 4655–4661
Truglio JJ, Croteau DL, Van Houten B, Kisker C (2006) Prokaryotic
nucleotide excision repair: the UvrABC system. Chem Rev 106:
233–252
Tsuzuki T, Fujii Y, Sakumi K, Tominaga Y, Nakao K, Sekiguchi M,
Matsushiro A, Yoshimura Y, Morita T (1996a) Targeted disruption
of the Rad51 gene leads to lethality in embryonic mice. Proc Natl
Acad Sci USA 93: 6236–6240
Tsuzuki T, Sakumi K, Shiraishi A, Kawate H, Igarashi H,
Iwakuma T, Tominaga Y, Zhang S, Shimizu S, Ishikawa T,
Nakamura K, Nakano K, Katsuki M, Sekiguchi M (1996b)
Targeted disruption of the DNA repair methyltransferase gene
renders mice hypersensitive to alkylating agent. Carcinogenesis
17: 1215–1220
van der Horst GT, Meira L, Gorgels TG, de Wit J, Velasco-Miguel S,
Richardson JA, Kamp Y, Vreeswijk MP, Smit B, Bootsma D,
Hoeijmakers JH, Friedberg EC (2002) UVB radiation-induced
cancer predisposition in Cockayne syndrome group A (Csa)
mutant mice. DNA Repair (Amst) 1: 143–157
van der Horst GT, van Steeg H, Berg RJ, van Gool AJ, de Wit J,
Weeda G, Morreau H, Beems RB, van Kreijl CF, de Gruijl FR,
Bootsma D, Hoeijmakers JH (1997) Defective transcription-
coupled repair in Cockayne syndrome B mice is associated with
skin cancer predisposition. Cell 89: 425–435
Vasen HF, Moslein G, Alonso A, Bernstein I, Bertario L, Blanco I,
Burn J, Capella G, Engel C, Frayling I, Friedl W, Hes FJ, Hodgson
S, Mecklin JP, Moller P, Nagengast F, Parc Y, Renkonen-Sinisalo L,
Sampson JR, Stormorken A et al (2007) Guidelines for the clinical
management of Lynch syndrome (hereditary non-polyposis
cancer). J Med Genet 44: 353–362
Veigl ML, Kasturi L, Olechnowicz J, Ma AH, Lutterbaugh JD,
Periyasamy S, Li GM, Drummond J, Modrich PL, Sedwick WD,
Markowitz SD (1998) Biallelic inactivation of hMLH1 by
epigenetic gene silencing, a novel mechanism causing human
MSI cancers. Proc Natl Acad Sci USA 95: 8698–8702
Weeda G, Donker I, de Wit J, Morreau H, Janssens R, Vissers CJ,
Nigg A, van Steeg H, Bootsma D, Hoeijmakers JH (1997) Disruption
of mouse ERCC1 results in a novel repair syndrome with growth
failure, nuclear abnormalities and senescence. Curr Biol 7: 427–439
Wei K, Clark AB, Wong E, Kane MF, Mazur DJ, Parris T, Kolas NK,
Russell R, Hou Jr H, Kneitz B, Yang G, Kunkel TA, Kolodner RD,
Cohen PE, Edelmann W (2003) Inactivation of exonuclease 1 in mice
results in DNA mismatch repair defects, increased cancer suscept-
ibility, and male and female sterility. Genes Dev 17: 603–614
Wesoly J, Agarwal S, Sigurdsson S, Bussen W, Van Komen S, Qin J,
van Steeg H, van Benthem J, Wassenaar E, Baarends WM,
Ghazvini M, Tafel AA, Heath H, Galjart N, Essers J, Grootegoed JA,
Arnheim N, Bezzubova O, Buerstedde JM, Sung P et al (2006)
Differential contributions of mammalian Rad54 paralogs to recombi-
nation, DNA damage repair, and meiosis. Mol Cell Biol 26: 976–989
West SC (1997) Processing of recombination intermediates by the
RuvABC proteins. Annu Rev Genet 31: 213–244
Wiesendanger M, Kneitz B, Edelmann W, Scharff MD
(2000) Somatic hypermutation in MutS homologue (MSH)3-,
MSH6-, and MSH3/MSH6-deficient mice reveals a role for the
MSH2–MSH6 heterodimer in modulating the base substitution
pattern. J Exp Med 191: 579–584
Williams BR, Mirzoeva OK, Morgan WF, Lin J, Dunnick W, Petrini
JH (2002) A murine model of Nijmegen breakage syndrome. Curr
Biol 12: 648–653
Wilson III DM, Bohr VA (2007) The mechanics of base excision
repair, and its relationship to aging and disease. DNA Repair
(Amst) 6: 544–559
Woo Y, Wright SM, Maas SA, Alley TL, Caddle LB, Kamdar S,
Affourtit J, Foreman O, Akeson EC, Shaffer D, Bronson RT, Morse
III HC, Roopenian D, Mills KD (2007) The nonhomologous end
joining factor Artemis suppresses multi-tissue tumor formation
and prevents loss of heterozygosity. Oncogene 26: 6010–6020
Xiao Y, Weaver DT (1997) Conditional gene targeted deletion by Cre
recombinase demonstrates the requirement for the double-strand
break repair Mre11 protein in murine embryonic stem cells.
Nucleic Acids Res 25: 2985–2991
Xie Y, Yang H, Cunanan C, Okamoto K, Shibata D, Pan J, Barnes DE,
Lindahl T, McIlhatton M, Fishel R, Miller JH (2004) Deficiencies
in mouse Myh and Ogg1 result in tumor predisposition and G to T
mutations in codon 12 of the K-ras oncogene in lung tumors.
Cancer Res 64: 3096–3102
Xu X, Qiao W, Linke SP, Cao L, Li WM, Furth PA, Harris CC, Deng CX
(2001) Genetic interactions between tumor suppressors Brca1 and
p53 in apoptosis, cell cycle and tumorigenesis. Nat Genet 28: 266–271
Xu X, Wagner KU, Larson D, Weaver Z, Li C, Ried T, Hennighausen
L, Wynshaw-Boris A, Deng CX (1999) Conditional mutation of
Brca1 in mammary epithelial cells results in blunted ductal
morphogenesis and tumour formation. Nat Genet 22: 37–43
Yamaguchi-Iwai Y, Sonoda E, Buerstedde JM, Bezzubova O,
Morrison C, Takata M, Shinohara A, Takeda S (1998)
Homologous recombination, but not DNA repair, is reduced in
vertebrate cells deficient in RAD52. Mol Cell Biol 18: 6430–6435
Yamaguchi-Iwai Y, Sonoda E, Sasaki MS, Morrison C, Haraguchi T,
Hiraoka Y, Yamashita YM, Yagi T, Takata M, Price C, Kakazu N,
Takeda S (1999) Mre11 is essential for the maintenance of
chromosomal DNA in vertebrate cells. EMBO J 18: 6619–6629
Yan CT, Kaushal D, Murphy M, Zhang Y, Datta A, Chen C, Monroe
B, Mostoslavsky G, Coakley K, Gao Y, Mills KD, Fazeli AP,
Tepsuporn S, Hall G, Mulligan R, Fox E, Bronson R, De
Girolami U, Lee C, Alt FW (2006) XRCC4 suppresses medullo-
blastomas with recurrent translocations in p53-deficient mice.
Proc Natl Acad Sci USA 103: 7378–7383
Yoon T, Chakrabortty A, Franks R, Valli T, Kiyokawa H,
Raychaudhuri P (2005) Tumor-prone phenotype of the DDB2-
deficient mice. Oncogene 24: 469–478
Yoshida K, Kondoh G, Matsuda Y, Habu T, Nishimune Y, Morita T
(1998) The mouse RecA-like gene Dmc1 is required for homo-
logous chromosome synapsis during meiosis. Mol Cell 1: 707–718
Zha S, Alt FW, Cheng HL, Brush JW, Li G (2007) Defective DNA
repair and increased genomic instability in Cernunnos–XLF-defi-
cient murine ES cells. Proc Natl Acad Sci USA 104: 4518–4523
Zhu J, Petersen S, Tessarollo L, Nussenzweig A (2001) Targeted
disruption of the Nijmegen breakage syndrome gene NBS1 leads
to early embryonic lethality in mice. Curr Biol 11: 105–109
DNA-damage repair
R Hakem
&2008 European Molecular Biology Organization The EMBO Journal VOL 27 |NO 4 |2008 605
... This damage occurs randomly across the DNA structure, overwhelming repair mechanisms. In line with current understanding, if DNA repair mechanisms fail, cells activate programmed cell death pathways, such as apoptosis [13]. Essentially, cells in a tissue that sustain DNA damage beyond repair capacity are eliminated from the population. ...
Article
Full-text available
Glioma tumors are the most common form of central nervous system tumors, and there is a pressing need for innovative methods that can precisely target cancer cells while leaving healthy tissues unharmed. In this study, progressing in the field of Catalytic Nanomedicine, we investigated the cytotoxic effects of a novel class of bionanocatalysts on glioma cancer cells. These bionanocatalysts were constructed from a catalytic matrix of oxides with evenly dispersed superficial copper-coating nanoparticles. This design optimizes both the inherent catalytic characteristics of the matrix and instills cytotoxic properties. The bionanocatalysts coated with copper demonstrated a significant reduction in cancer cell viability when compared to reference bionanocatalysts without the transition metal. We also observed structural damage to the cytoskeleton and alterations in mitochondrial activity. These findings suggest that these pathways are integral to the mechanisms through which these nanostructures execute their bionanocatalytic activities, particularly in breaking chemical bonds. Importantly, our physicochemical analyses verified that the coating with copper species, primarily CuO, did not disrupt the individual structure of the bionanocatalysts: instead, it enhanced their catalytic cytotoxic potential. This research aims to deepen our understanding of the mechanisms underlying this promising antineoplastic treatment and underscore the effectiveness of superficial copper-coating nanoparticles as agents for amplifying the inherent properties of bionanocatalysts through nanocatalysis.
... DNA damage is one of the most severe consequences of Cd-induced oxidative stress, where the generated ROS not only cause base oxidation but also lead to DNA strand breaks and cross-linking. If such damage is not promptly and effectively repaired, it can result in gene mutations, chromosomal aberrations, and even cellular transformation into cancer [37]. Chronic exposure is associated with an increased risk of various cancers, partly due to the sustained DNA damage caused by increased intracellular ROS production. ...
Article
Full-text available
Cadmium (Cd), a prevalent environmental contaminant, exerts widespread toxic effects on human health through various biochemical and molecular mechanisms. This review encapsulates the primary pathways through which Cd inflicts damage, including oxidative stress induction, disruption of Ca2+ signaling, interference with cellular signaling pathways, and epigenetic modifications. By detailing the absorption, distribution, metabolism, and excretion (ADME) of Cd, alongside its interactions with cellular components such as mitochondria and DNA, this paper highlights the extensive damage caused by Cd2+ at the cellular and tissue levels. The role of Cd in inducing oxidative stress—a pivotal mechanism behind its toxicity—is discussed with emphasis on how it disrupts the balance between oxidants and antioxidants, leading to cellular damage and apoptosis. Additionally, the review covers Cd’s impact on signaling pathways like Mitogen-Activated Protein Kinase (MAPK), Nuclear Factor kappa-light-chain-enhancer of activated B cells (NF-κB), and Tumor Protein 53 (p53) pathways, illustrating how its interference with these pathways contributes to pathological conditions and carcinogenesis. The epigenetic effects of Cd, including DNA methylation and histone modifications, are also explored to explain its long-term impact on gene expression and disease manifestation. This comprehensive analysis not only elucidates the mechanisms of Cd toxicity but also underscores the critical need for enhanced strategies to mitigate its public health implications.
... 87 Other interesting interacting partners involved in DNA repair are BRCA2, RAD21, LIG1, TRP53, and NHEJ1. 88 Our mass spectrometry data revealed strong enrichment for MSH2 in all 4 replicates of the IP using anti-TET3 over the IgG control (Table S1). MSH2 has been found to interact with TET2 in ESCs 43 (Table S2). ...
Article
Full-text available
Ten-eleven translocation (TET) proteins are DNA dioxygenases that mediate active DNA demethylation. TET3 is the most highly expressed TET protein in thymic developing T cells. TET3, either independently or in cooperation with TET1 or TET2, has been implicated in T cell lineage specification by regulating DNA demethylation. However, TET-deficient mice exhibit complex phenotypes, suggesting that TET3 exerts multifaceted roles, potentially by interacting with other proteins. We performed liquid chromatography with tandem mass spectrometry in primary developing T cells to identify TET3 interacting partners in endogenous, in vivo conditions. We discover TET3 interacting partners. Our data establish that TET3 participates in a plethora of fundamental biological processes, such as transcriptional regulation, RNA polymerase elongation, splicing, DNA repair, and DNA replication. This resource brings in the spotlight emerging functions of TET3 and sets the stage for systematic studies to dissect the precise mechanistic contributions of TET3 in shaping T cell biology.
... Alterations to DNA sequence or structure (at the nucleotide level) are repaired by three main pathways: nucleotide excision repair (NER), base excision repair (BER), or mismatch repair (MMR). These pathways are activated in response to bulky DNA adducts formed due to UV radiation or chemical carcinogens, damaged bases due to spontaneous oxidation, alkylation, or deamination, or mismatched bases due to errors in DNA replication, respectively (1). Interestingly, viruses seldom hijack components of these pathways. ...
Article
Full-text available
Viruses are obligate intracellular pathogens that hijack a myriad of host cell processes to facilitate replication and suppress host antiviral defenses. In its essence, a virus is a segment of foreign nucleic acid that engages host cell machinery to drive viral genome replication, gene transcription, and protein synthesis to generate progeny virions. Because of this, host organisms have developed sophisticated detection systems that activate antiviral defenses following recognition of aberrant nucleic acids. For example, recognition of viral nucleic acids by host DNA repair proteins results in compromised viral genome integrity, induction of antiviral inflammatory programs, cell cycle arrest, and apoptosis. Unsurprisingly, diverse viral families have evolved multiple strategies that fine-tune host DNA repair responses to suppress activation of antiviral defenses while simultaneously hijacking DNA repair proteins to facilitate virus replication. This review summarizes common molecular strategies viruses deploy to exploit host DNA repair mechanisms.
... Lipid peroxidation refers to the oxidative degradation of lipids, which can have detrimental effects on cell membranes and overall cell function. Apoptosis induced by Sudan I's breakdown products implies that the dye has the potential to trigger programmed cell death, a crucial mechanism to eliminate cells with irreparable DNA damage or other significant abnormalities (Hakem 2008). While the existing evidence highlights the genotoxic and carcinogenic potential of Sudan I dye, it is essential to emphasize the need for more in-depth in vivo investigations to fully understand the complexities of its toxic profile. ...
Article
Full-text available
The textile industry, with its extensive use of dyes and chemicals, stands out as a significant source of water pollution. Exposure to certain textile dyes, such as azo dyes and their breakdown products like aromatic amines, has been associated with health concerns like skin sensitization, allergic reactions, and even cancer in humans. Annually, the worldwide production of synthetic dyes approximates 7 × 10⁷ tons, of which the textile industry accounts for over 10,000 tons. Inefficient dyeing procedures result in the discharge of 15–50% of azo dyes, which do not adequately bind to fibers, into wastewater. This review delves into the genotoxic impact of azo dyes, prevalent in the textile industry, on aquatic ecosystems and human health. Examining different families of textile dye which contain azo group in their structure such as Sudan I and Sudan III Sudan IV, Basic Red 51, Basic Violet 14, Disperse Yellow 7, Congo Red, Acid Red 26, and Acid Blue 113 reveals their carcinogenic potential, which may affect both industrial workers and aquatic life. Genotoxic and carcinogenic characteristics, chromosomal abnormalities, induced physiological and neurobehavioral changes, and disruptions to spermatogenesis are evident, underscoring the harmful effects of these dyes. The review calls for comprehensive investigations into the toxic profile of azo dyes, providing essential insights to safeguard the aquatic ecosystem and human well-being. The importance of effective effluent treatment systems is underscored to mitigate adverse impacts on agricultural lands, water resources, and the environment, particularly in regions heavily reliant on wastewater irrigation for food production.
Article
Error correction is central to many biological systems and is critical for protein function and cell health. During mitosis, error correction is required for the faithful inheritance of genetic material. When functioning properly, the mitotic spindle segregates an equal number of chromosomes to daughter cells with high fidelity. Over the course of spindle assembly, many initially erroneous attachments between kinetochores and microtubules are fixed through the process of error correction. Despite the importance of chromosome segregation errors in cancer and other diseases, there is a lack of methods to characterize the dynamics of error correction and how it can go wrong. Here, we present an experimental method and analysis framework to quantify chromosome segregation error correction in human tissue culture cells with live cell confocal imaging, timed premature anaphase, and automated counting of kinetochores after cell division. We find that errors decrease exponentially over time during spindle assembly. A coarse-grained model, in which errors are corrected in a chromosome-autonomous manner at a constant rate, can quantitatively explain both the measured error correction dynamics and the distribution of anaphase onset times. We further validated our model using perturbations that destabilized microtubules and changed the initial configuration of chromosomal attachments. Taken together, this work provides a quantitative framework for understanding the dynamics of mitotic error correction.
Article
Cancer care has witnessed remarkable progress in recent decades, with a wide array of targeted therapies and immune-based interventions being added to the traditional treatment options such as surgery, chemotherapy, and radiotherapy. However, despite these advancements, the challenge of achieving high tumor specificity while minimizing adverse side effects continues to dictate the benefit-risk balance of cancer therapy, guiding clinical decision making. As such, the targeting of cancer testis antigens (CTAs) offers exciting new opportunities for therapeutic intervention of cancer since they display highly tumor specific expression patterns, natural immunogenicity and play pivotal roles in various biological processes that are critical for tumor cellular fitness. In this review, we delve deeper into how CTAs contribute to the regulation and maintenance of genomic integrity in cancer, and how these mechanisms can be exploited to specifically target and eradicate tumor cells. We review the current clinical trials targeting aforementioned CTAs, highlight promising pre-clinical data and discuss current challenges and future perspectives for future development of CTA-based strategies that exploit tumor genomic instability.
Article
Full-text available
Background The remarkable resistance to ionizing radiation found in anhydrobiotic organisms, such as some bacteria, tardigrades, and bdelloid rotifers has been hypothesized to be incidental to their desiccation resistance. Both stresses produce reactive oxygen species and cause damage to DNA and other macromolecules. However, this hypothesis has only been investigated in a few species. Results In this study, we analyzed the transcriptomic response of the bdelloid rotifer Adineta vaga to desiccation and to low- (X-rays) and high- (Fe) LET radiation to highlight the molecular and genetic mechanisms triggered by both stresses. We identified numerous genes encoding antioxidants, but also chaperones, that are constitutively highly expressed, which may contribute to the protection of proteins against oxidative stress during desiccation and ionizing radiation. We also detected a transcriptomic response common to desiccation and ionizing radiation with the over-expression of genes mainly involved in DNA repair and protein modifications but also genes with unknown functions that were bdelloid-specific. A distinct transcriptomic response specific to rehydration was also found, with the over-expression of genes mainly encoding Late Embryogenesis Abundant proteins, specific heat shock proteins, and glucose repressive proteins. Conclusions These results suggest that the extreme resistance of bdelloid rotifers to radiation might indeed be a consequence of their capacity to resist complete desiccation. This study paves the way to functional genetic experiments on A. vaga targeting promising candidate proteins playing central roles in radiation and desiccation resistance.
Article
Full-text available
Breast cancer is a chief cause of cancer-related mortality that affects women worldwide. About 8% of cases are hereditary, and approximately half of these are associated with germline mutations of the breast tumor suppressor gene BRCA1 (refs. 1,2). We have previously reported a mouse model in which Brca1 exon 11 is eliminated in mammary epithelial cells through Cre-mediated excision. This mutation is often accompanied by alterations in transformation-related protein 53 (Trp53, encoding p53), which substantially accelerates mammary tumor formation. Here, we sought to elucidate the underlying mechanism(s) using mice deficient in the Brca1 exon 11 isoform (Brca1Delta11/Delta11). Brca1Delta11/Delta11 embryos died late in gestation because of widespread apoptosis. Unexpectedly, elimination of one Trp53 allele completely rescues this embryonic lethality and restores normal mammary gland development. However, most female Brca1Delta11/Delta11 Trp53+/- mice develop mammary tumors with loss of the remaining Trp53 allele within 6-12 months. Lymphoma and ovarian tumors also occur at lower frequencies. Heterozygous mutation of Trp53 decreases p53 and results in attenuated apoptosis and G1-S checkpoint control, allowing Brca1Delta11/Delta11 cells to proliferate. The p53 protein regulates Brca1 transcription both in vitro and in vivo, and Brca1 participates in p53 accumulation after gamma-irradiation through regulation of its phosphorylation and Mdm2 expression. These findings provide a mechanism for BRCA1-associated breast carcinogenesis.
Article
Full-text available
textabstractWe have generated mice deficient in O6-methylguanine DNA methyltransferase activity encoded by the murine Mgmt gene using homologous recombination to delete the region encoding the Mgmt active site cysteine. Tissues from Mgmt null mice displayed very low O6-methylguanine DNA methyltransferase activity, suggesting that Mgmt constitutes the major, if not the only, O6-methylguanine DNA methyltransferase. Primary mouse embryo fibroblasts and bone marrow cells from Mgmt -/- mice were significantly more sensitive to the toxic effects of the chemotherapeutic alkylating agents 1,3-bis(2-chloroethyl)-1-nitrosourea, streptozotocin and temozolomide than those from Mgmt wild-type mice. As expected, Mgmt-deficient fibroblasts and bone marrow cells were not sensitive to UV light or to the crosslinking agent mitomycin C. In addition, the 50% lethal doses for Mgmt -/- mice were 2- to 10-fold lower than those for Mgmt +/+ mice for 1,3-bis(2chloroethyl)-1-nitrosourea, N-methyl-N-nitrosourea and streptozotocin; similar 50% lethal doses were observed for mitomycin C. Necropsies of both wild-type and Mgmt -/mice following drug treatment revealed histological evidence of significant ablation of hematopoietic tissues, but such ablation occurred at much lower doses for the Mgmt -/- mice. These results demonstrate the critical importance of O6-methylguanine DNA methyltransferase in protecting cells and animals against the toxic effects of alkylating agents used for cancer chemotherapy.
Article
Full-text available
Repair of mismatches in DNA in mammalian cells is mediated by a complex of proteins that are members of two highly conserved families of genes referred to as MutS and MutL homologues. Germline mutations in several members of these families, MSH2, MSH6, MLH1, and PMS2, but not MSH3, are responsible for hereditary non-polyposis colorectal cancer. To examine the role of MSH3, we generated a mouse with a null mutation in this gene. Cells from Msh32/2 mice are defective in repair of insertion/ deletion mismatches but can repair base-base mismatches. Msh32/2 mice develop tumors at a late age. When the Msh32/2 and Msh62/2 muta- tions are combined, the tumor predisposition phenotype is indistinguish- able from Msh22/2 or Mlh12/2 mice. These results suggest that MSH3 cooperates with MSH6 in tumor suppression.
Article
By removing biosynthetic errors from newly synthesized DNA, mismatch repair (MMR) improves the fidelity of DNA replication by several orders of magnitude. Loss of MMR brings about a mutator phenotype, which causes a predisposition to cancer. But MMR status also affects meiotic and mitotic recombination, DNA-damage signalling, apoptosis and cell-type-specific processes such as class-switch recombination, somatic hypermutation and triplet-repeat expansion. This article reviews our current understanding of this multifaceted DNA-repair system in human cells.
Article
The enzyme O 6 -methylguanine-DNA methyltransferase repairs alkylation-induced DNA damage, O 6 -methylguan-ine and O 4 -methylthymine, the former being formed more frequently. Previously, by means of gene targeting, we generated mice in which alleles for methyltransferase were disrupted. We now use these mouse lines, which are totally deficient in methyltransferase activity, to examine protective effects of the enzyme against tumor formation. In gene-targeted female mice given an i.p. injection of 5 mg/kg of dimethylnitrosamine, a larger number of liver and lung tumors occurred, as compared with normal female mice treated in the same manner. In male mice given a lower dose of carcinogen, the difference between normal and gene-targeted mice was statistically insignificant although more tumors did form in the gene-targeted mice. Methyltransferase apparently afforded protection from nitrosamine-induced tumorigenesis.
Article
Comparison of the clinical and cellular phenotypes of different genomic instability syndromes provides new insights into functional links in the complex network of the DNA damage response. A prominent example of this principle is provided by examination of three such disorders: ataxia-telangiectasia (A-T) caused by lack or inactivation of the ATM protein kinase, which mobilises the cellular response to double strand breaks in the DNA; ataxia-telangiectasia-like disease (ATLD), a result of deficiency of the human Mre11 protein; and the Nijmegen breakage syndrome (NBS), which represents defective Nbs1 protein. Mre11 and Nbs1 are members of the Mre11/Rad50/Nbs1 (MRN) protein complex. MRN and its individual components are involved in different responses to cellular damage induced by ionising radiation and radiomimetic chemicals, including complexing with chromatin and with other damage response proteins, formation of radiation-induced foci, and the induction of different cell cycle checkpoints. The phosphorylation of Nbs1 by ATM would indicate that ATM acts upstream of the MRN complex. Consistent with this were the suggestions that ATM could be activated in the absence of fully functional Nbs1 protein. In contrast, the regulation of some ATM target proteins, e.g. Smc1 requires the MRN complex as well as ATM. Nbs1 may, therefore, be both a substrate for ATM and a mediator of ATM function. Recent studies that indicate a requirement of the MRN complex for proper ATM activation suggest that the relationship between ATM and the MRN complex in the DNA damage response is yet to be fully determined. Despite the fact that both Mre11 and Nbs1 are part of the same MRN complex, deficiency in either protein in humans does not lead to the same clinical picture. This suggests that components of the complex may also act separately.
Article
Nitrosoureas form O6-alkylguanine-DNA adducts that are converted to G to A transitions, the mutation found in the activated ras oncogenes of nitrosourea-induced mouse lymphomas and rat mammary tumors. These adducts are removed by the DNA repair protein O6-alkylguanine-DNA alkyltransferase. Transgenic mice that express the human homolog of this protein in the thymus were found to be protected from developing thymic lymphomas after exposure to N-methyl-N-nitrosourea. Thus, transgenic expression of a single human DNA repair gene is sufficient to block chemical carcinogenesis. The transduction of DNA repair genes in vivo may unravel mechanisms of carcinogenesis and provide therapeutic protection from known carcinogens.