Download The role of DNA damage in laminopathy progeroid syndromes

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

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

Document related concepts

Comparative genomic hybridization wikipedia , lookup

Genomic library wikipedia , lookup

Telomere wikipedia , lookup

Mitochondrial DNA wikipedia , lookup

Epigenetics of diabetes Type 2 wikipedia , lookup

DNA methylation wikipedia , lookup

DNA profiling wikipedia , lookup

SNP genotyping wikipedia , lookup

Zinc finger nuclease wikipedia , lookup

DNA polymerase wikipedia , lookup

No-SCAR (Scarless Cas9 Assisted Recombineering) Genome Editing wikipedia , lookup

Microevolution wikipedia , lookup

Oncogenomics wikipedia , lookup

Mutation wikipedia , lookup

Genomics wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

Replisome wikipedia , lookup

Polycomb Group Proteins and Cancer wikipedia , lookup

Epigenetics wikipedia , lookup

Epigenetics in stem-cell differentiation wikipedia , lookup

Epigenetics of neurodegenerative diseases wikipedia , lookup

Mutagen wikipedia , lookup

Gel electrophoresis of nucleic acids wikipedia , lookup

Primary transcript wikipedia , lookup

Genealogical DNA test wikipedia , lookup

Epigenetics in learning and memory wikipedia , lookup

United Kingdom National DNA Database wikipedia , lookup

Nucleic acid analogue wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Epigenetic clock wikipedia , lookup

Genome editing wikipedia , lookup

DNA repair wikipedia , lookup

Molecular cloning wikipedia , lookup

Non-coding DNA wikipedia , lookup

Nutriepigenomics wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

DNA vaccination wikipedia , lookup

Bisulfite sequencing wikipedia , lookup

Point mutation wikipedia , lookup

Cell-free fetal DNA wikipedia , lookup

Nucleic acid double helix wikipedia , lookup

Nucleosome wikipedia , lookup

DNA supercoil wikipedia , lookup

Helitron (biology) wikipedia , lookup

History of genetic engineering wikipedia , lookup

Cre-Lox recombination wikipedia , lookup

Extrachromosomal DNA wikipedia , lookup

Deoxyribozyme wikipedia , lookup

Cancer epigenetics wikipedia , lookup

Epigenomics wikipedia , lookup

DNA damage theory of aging wikipedia , lookup

Transcript
Nuclear Envelope Disease and Chromatin Organization
The role of DNA damage in laminopathy progeroid
syndromes
Christopher J. Hutchison1
School of Biological and Biomedical Sciences, Durham University, South Road, Durham DH1 3LE, U.K.
Abstract
Progeroid laminopathies are characterized by the abnormal processing of lamin A, the appearance of
misshapen nuclei, and the accumulation and persistence of DNA damage. In the present article, I consider
the contribution of defective DNA damage pathways to the pathology of progeroid laminopathies. Defects
in DNA repair pathways appear to be caused by a combination of factors. These include abnormal epigenetic
modifications of chromatin that are required to recruit DNA repair pathways to sites of DNA damage,
abnormal recruitment of DNA excision repair proteins to sites of DNA double-strand breaks, and unrepairable
ROS (reactive oxygen species)-induced DNA damage. At least two of these defective processes offer the
potential for novel therapeutic approaches.
The progeroid laminopathies and prelamin
A toxicity
The progeroid laminopathies include HGPS (Hutchinson–
Gilford progeria syndrome), atypical Werner’s syndrome,
MAD (mandibuloacral dysplasia) and RD (restrictive
dermopathy) [1]. All of these diseases have been linked to
abnormal post-translational processing of prelamin A, which
promotes nuclear shape abnormalities and cellular toxicity
[2–5]. The initial post-translational processing of lamin A
is now well-characterized. Lamin A is first translated as a
precursor molecule termed prelamin A that possesses a Cterminal motif, CaaX (where a is an aliphatic residue), which
is a receptor site for protein farnesylation and methylation.
Initially, the cysteine residue is farnesylated before the final
three amino acids are proteolytically cleaved. This is followed
by methylation of the now C-terminal cysteine residue before
the partially processed prelamin A is translocated to the
INM (inner nuclear membrane). Once at the INM,
the final 15 amino acids are cleaved by the ZmpSte24
(zinc metallopeptidase Ste24 homologue) protease to generate
mature lamin A [3,6]. A majority (∼ 80%) of mutations giving
rise to HGPS are caused by silent mutations in exon 11
of LMNA that activates a cryptic splice site and permits
synthesis of an alternatively splice variant of prelamin A,
termed progerin, which lacks the ZmpSte24 cleavage site and
consequently cannot undergo the final processing step. This
in turn means that progerin is permanently farnesylated [7,8].
The representation of progerin in fibroblasts from HGPS
Key words: DNA damage response, Hutchinson–Gilford progeria syndrome (HGPS), lamin,
laminopathy, progeria, reactive oxygen species (ROS).
Abbreviations used: ATM, ataxia telangiectasia mutated; ATR, ATM- and Rad3-related; 53BP1,
p53-binding protein 1; DDR, DNA damage response; DSB, double-strand break; FC-prelamin
A, prelamin A with farnesylated cysteine residue; FTI, farnesyltransferase inhibitor; HGPS,
Hutchinson–Gilford progeria syndrome; HP1, heterochromatin protein 1; INM, inner nuclear
membrane; MAD, mandibuloacral dysplasia; MEF, mouse embryonic fibroblast; NAC, Nacetylcysteine; NuRD, nucleosome remodelling and deacetylation; RBBP, retinoblastoma-binding
protein; RD, restrictive dermopathy; ROS, reactive oxygen species; siRNA, small interfering RNA;
XPA, xeroderma pigmentosum group A; ZmpSte24, zinc metallopeptidase Ste24 homologue.
1
email [email protected]
Biochem. Soc. Trans. (2011) 39, 1715–1718; doi:10.1042/BST20110700
patients increases with increasing passage number [9] and
this is thought to increase its toxic effects. In RD and MAD,
as well as in some cases of HGPS, the diseases are caused by
homozygous mutations in a conserved region of the gene
encoding the ZmpSte24 protease, and in fibroblasts from
these patients, the disease is caused by abnormal processing of
prelamin A, leading to the accumulation of a form of prelamin
A that retains the farnesylated cysteine residue (FCprelamin A) [10–12]. Some forms of HGPS (non-classical
HGPS) as well as atypical Werner’s syndrome are caused by
mutations in exons of LMNA encoding the rod domain of
lamins A and C [1] and do not give rise to splice site variants,
whereas other mutations occur in the BANF1 gene [encoding
the chromatin protein BAF (barrier to autointegration factor
1)] [13], suggesting that laminopathy progerias do not
necessarily all arise through the accumulation of farnesylated
forms of lamin A.
Despite reports of non-classical forms of HGPS, therapeutic approaches have centred on the use of drugs that inhibit
the farnesylation of both progerin and prelamin A. These
include the use of FTIs (farnesyltransferase inhibitors) as a
direct tool to inhibit farnesylation [2,4,5] or a combination
of aminobisphosphonates and statins, which act indirectly by
preventing the synthesis of precursors of farnesyl and geranyl
residues [14]. Both types of treatment do extend lifespan in
mouse models of HGPS [14] and are currently being used
in clinical trials.
DDR (DNA damage response) in progeroid
laminopathies
As stated above, the most obvious morphological feature of
the progeroid laminopathies is the appearance of misshapen
(or dysmorphic) nuclei in cells obtained from patients or in
cells that express progerin ectopically [9]. The presence of
dysmorphic nuclei can be ameliorated by growing cells in
media containing FTIs [2,4,5]. Another feature of fibroblasts
C The
C 2011 Biochemical Society
Authors Journal compilation 1715
1716
Biochemical Society Transactions (2011) Volume 39, part 6
from progeroid laminopathy patients is the accumulation of
unrepaired DNA damage [15–17] and accelerated telomere
attrition [18,19]. The accumulation of unrepaired DNA
damage activates a checkpoint response that is characterized
by phosphorylation of both the ATM (ataxia telangiectasia
mutated) and ATR (ATM- and Rad3-related) checkpoint
kinases. Activation of checkpoint responses seems to underlie
the slow growth of HGPS fibroblasts in culture, since siRNA
(small interfering RNA) knockdown of ATM and ATR in
these cells restores replicative capacity [20]. In stark contrast
with nuclear shape abnormalities, accumulation of unrepaired
DNA damage is not reversed by treatment with FTIs [20]
and this finding is consistent with studies showing that the
accumulation of unfarnesylated forms of progerin in mice
still promotes aging phenotypes [21], whereas treatment of
Caenorhabditis elegans with FTIs limits the accumulation
of dysmophic nuclei, without extending lifespan [22]. These
findings argue strongly that processes other those promoting
the formation of dysmorphic nuclei account for the aging
phenotypes in laminopathy progerias and has led to a
number of investigations into the causes and effects of the
accumulation of unrepairable DNA damage in these diseases.
The first study revealing the increased sensitivity to
ionizing radiation and genome instability in a mouse model
of HGPS and fibroblasts from HGPS patients demonstrated
that the DDR was impaired and that this was characterized
by delayed recruitment of DNA damage proteins, including
53BP1 (p53-binding protein 1) and Rad50 to sites of
DNA damage [15]. The impaired DDR was correlated with
activation of p53-response pathways, although, curiously,
this occurred in the absence of increased levels of p53 or p53
phosphorylation [23]. DDRs are characterized by epigenetic
modification of chromosomes, including the accumulation
of the phosphorylated form of a histone H2A variant
H2AX at sites of DNA DSBs (double-strand breaks) [24]
and methylation of histone H3 and histone H4 [25,26].
A hallmark of fibroblasts from progeriod laminopathy
patients is the persistence of H2AX foci [15], altered
methylation of H3 and H4 and down-regulation of HP1
(heterochromatin protein 1) [27,28], which are all predicted
to lead to a loss of heterochromatin. The persistence of
H2AX foci is indicative of an accumulation of unrepaired
DSBs. However, there is now strong evidence that this
is underpinned by altered H3 methylation. The NuRD
(nucleosome remodelling and deacetylation) complex is a key
regulator of histone methylation [29]. Chromatin proteins
RBBP (retinoblastoma-binding protein) 4 and its paralogue RBBP7 are both components of the NuRD complex
and bind to the region of the lamin A tail domain, which
is lacking in progerin. The levels of both RBBP4 and
RBBP7 are reduced in HGPS fibroblasts compared with agematched control fibroblasts, and this loss is dependent upon
the presence of progerin. Interestingly, loss of RBBP4 and
RBBP7 precedes the accumulation of persistent H2AX foci,
and siRNA knockdown of RBBP4 and RBBP7 in normal
fibroblasts leads to decreased H3 methylation followed
by the accumulation of H2AX foci. Knockdown of other
C The
C 2011 Biochemical Society
Authors Journal compilation components of the NuRD complex also promote the
accumulation of unrepaired DSBs [30]. In a more recent
study, hypoacetylation of H4 was observed in fibroblasts
from a ZmpSte24 − / − mouse. Hypoacetylation of H4 was
linked to loss of nuclear matrix association of the histone
acetyltransferase Mof. Importantly, overexpression of Mof
in ZmpSte24 − / − fibroblasts or treatment of the cells with
a histone deacetylase inhibitor promoted the recruitment
of DNA repair proteins to sites of DNA damage and
significantly decreased the rate of entry into a senescent
state [31]. Both of these studies therefore imply that the
impairment of the DDR, in the presence of either progerin or
FC-prelamin A, arises from defects in chromatin-remodelling
steps that precede the recruitment of DNA damage proteins
to sites of DNA repair.
Impairment of DDR in fibroblasts from HGPS and RD
patients appears to be also related to incorrect recognition of
types of DNA damage. As stated above, DSBs induced by
ionizing radiation in HGPS fibroblasts are characterized
by impaired recruitment of the DNA repair proteins
Rad50 and Rad51 [31]. Both HGPS and RD fibroblasts
accumulate DSBs during passage in culture, even in the
absence of mutagenesis, and therefore carry a mutational
load. Interestingly, when treated with genotoxic agents such
as camptothecin or etoposide, newly induced DSBs can be
repaired efficiently in both HGPS and RD fibroblasts [32,33].
In contrast, the accumulated load of DNA damage remains
unrepaired. These sites fail to recruit Rad50 and Rad51,
but instead recruit the DNA excision repair protein XPA
(xeroderma pigmentosum group A). Importantly, siRNA
knockdown of XPA in HGPS and RD fibroblasts partially
restores the ability of these cells to undergo DNA repair,
implying that the presence of XPA at sites of DSBs is involved
in the failure to recruit DSB proteins to those same sites [32].
In a separate study, loss of stability of the DNA repair
protein 53BP1 has been implicated in telomere erosion and
persistent DSBs at telomeric DNA. This study utilized MEFs
(mouse embryonic fibroblasts) from an Lmna − / − mouse. In
these MEFs, telomeric DNA was abnormally distributed in
the interphase nucleus being clustered towards the nuclear
periphery (although not located at the nuclear envelope)
compared with wild-type MEFs, in which telomeres
were distributed throughout the nucleus. This abnormal
distribution of telomeric DNA was correlated with a modest,
but significant, shortening of telomere length, but highly
significantly impaired H4 methylation and accumulation
of H2AX in telomeric chromatin [34]. Telomeric DNA
damage is repaired by NHEJ (non-homologous end-joining),
which is mediated by the DNA repair protein 53BP1 [35].
Consistent, with this finding, the overall levels of 53BP1 were
reduced in Lmna − / − MEFs [34].
ROS (reactive oxygen species) and
persistent DNA damage in HGPS and RD
Recent studies have reported on the increased sensitivity of
fibroblasts from progeria patients to oxidative stress [36].
Nuclear Envelope Disease and Chromatin Organization
Moreover, progerin has also been reported to promote the
accumulation of oxidized proteins in fibroblasts [37]. ROS
generation and imbalance is a major cause of DNA damage
[38] and could underlie the accumulation and persistence of
DSBs in HGPS and RD fibroblasts with increasing age in
culture. With this in mind, ROS levels and ROS sensitivity in
fibroblasts from HGPS and RD patients were investigated.
Significantly higher basal levels of ROS, as well as increased
sensitivity to ROS stimuli, were reported. Both of these
features could be ameliorated with a commonly used ROS
scavenger NAC (N-acetylcysteine). ROS stimulation led
to DNA damage, which was only partially repaired and
therefore increased the load of DSBs in HGPS and RD
fibroblasts. Interestingly, treatment of cells with NAC, before
ROS stimulation, permitted repair of all DSBs. Moreover,
growth of HGPS and RD fibroblasts in the presence of
NAC also significantly reduced the basal levels of DSBs. This
finding implies that the elevated basal levels of ROS in these
cells is the cause of the persistent DNA damage [33].
Links to normal aging
ROS generation and the accumulation of persistent DNA
damage have both been linked to cellular and organismal
aging [39,40]. Expression of progerin and prelamin A have
also been linked to cellular senescence and aging in normal
individuals [27,41]. It is therefore important to understand
whether the mechanisms underlying the impaired DDR
in laminopathy progerias also underlie the accumulation
of persistent DNA damage in normal aging. Components of
the NuRD complex, including RBBP4, RBBP7 and HP1, are
all depleted in fibroblasts from very old individuals relative
to fibroblasts from very young individuals, implying that at
least one of the chromatin-remodelling pathways necessary
for efficient DNA repair loses efficiency with increasing age
[30]. In contrast, fibroblasts from very old individuals do
not display increased basal levels of ROS or increased ROS
sensitivity when compared with fibroblasts from very young
individuals. Moreover, repair of ROS-induced DSBs is very
efficient in the same cells [33]. Thus the DNA repair defects in
laminopathy progeroid syndromes appear to be only a partial
phenocopy of normal aging.
New therapeutic targets
At least two of the DNA damage defects in progeroid laminopathies appear to offer novel targets for pharmacological
intervention. At present, the most successful therapies for
existing patients is a combination of aminobisphosphonates
and statins [14]. However, ZmpSte24 − / − mice that were
provided with a histone deacetylase inhibitor in their drinking
water displayed weight gain, increased bone density and
increased lifespan compared with untreated littermates [31].
Similarly, when grown in the presence of NAC, fibroblasts
from HGPS and RD patients displayed normal growth
kinetics and increased longevity [33]. Thus a combination of
therapeutic approaches, which includes targets in DNA repair
pathways, might offer significant improvements in quality of
life for patients with progeroid laminopathies.
Acknowledgements
I thank Shane Richards, Joanne Muter and Ian Gibbs-Seymour for
their contribution to work on DNA repair pathways.
Funding
Work in my laboratory is supported by the Association for
International Cancer Research, the JGW Patterson Foundation, the
Wellcome Trust and ONE NE.
References
1 Broers, J.V.L., Ramaekers, F.C.S., Bonne, G., Ben Yaou, R. and Hutchison,
C.J. (2006) Nuclear lamins: laminopathies and their role in premature
ageing. Physiol. Rev. 86, 967–1008
2 Capell, B.C., Erdos, M.R., Madigan, J.P., Fiordalis, J.J., Varga, R., Connelly,
K.N., Gordon, L.B., Der, C.J., Cox, A.D. and Collins, FS. (2005) Inhibiting
farnesylation of progerin prevents the characteristic nuclear blebbing of
Hutchinson–Gilford progeria syndrome. Proc. Natl. Acad. Sci. U.S.A. 102,
12879–12884
3 Corrigan, D.P., Kuszczak, D., Rusinol, A.E., Thewke, D.P., Hrycyna, C.A.,
Michaelis, S. and Sinensky, M.S. (2005) Prelamin A endoproteolytic
processing in vitro by recombinant Zmpste24. Biochem. J. 387, 129–138
4 Mallampalli, M.P., Huyer, G., Bendale, P., Gelb, M.H. and Michaelis, S.
(2005) Inhibiting farnesylation reverses the nuclear morphology defect
in a HeLa cell model for Hutchinson–Gilford progeria syndrome. Proc.
Natl. Acad. Sci. U.S.A. 102, 14416–14421
5 Toth, J.I., Yang, S.H., Qiao, X., Beigneux, A.P., Gelb, M.H., Moulson, C.L.,
Miner, J.H., Young, S.G. and Fong, L.G. (2005) Blocking protein
farnesyltransferase improves nuclear shape in fibroblasts from humans
with progeroid syndromes. Proc. Natl. Acad. Sci. U.S.A. 102,
12873–12878
6 Sinensky, M. (2000) Recent advances in the study of prenylated
proteins. Biochim. Biophys. Acta 1484, 93–106
7 Eriksson, M., Brown, W.T., Gordon, L.B., Glynn, M.W., Singer, J., Scott, L.,
Erdos, M.R., Robbins, C.M., Moses, T.Y., Berglund, P. et al. (2003)
Recurrent de novo point mutations in lamin A cause Hutchinson–Gilford
progeria syndrome. Nature 423, 293–298
8 De Sandre-Giovannoli, A., Bernard, R., Cau, P., Navarro, C., Amiel, J.,
Boccaccio, I., Lyonnet, S., Stewart, C.L., Munnich, A., Le Merrer, M. and
Lévy, N. (2003) Lamin A truncation in Hutchinson–Gilford progeria.
Science 300, 2055–1057
9 Goldman, R.D., Shumaker, D.K., Erdos, M.R., Eriksson, M., Goldman, A.E.,
Gordon, L.B., Gruenbaum, Y., Khuon, S., Mendez, M., Varga, R. and
Collins, F.S. (2004) Accumulation of mutant lamin A causes progressive
changes in nuclear architecture in Hutchinson–Gilford progeria syndrome.
Proc. Natl. Acad. Sci. U.S.A. 101, 8963–8968
10 Denecke, J., Brune, T., Feldhaus, T., Robenek, H., Kranz, C., Auchus, R.J.,
Agarwal, A.K. and Marquardt, T. (2006) A homozygous ZMPSTE24 null
mutation in combination with a heterozygous mutation in the LMNA
gene causes Hutchinson–Gilford progeria syndrome (HGPS): insights into
the pathophysiology of HGPS. Hum. Mutat. 27, 524–531
11 Navarro, C.L., Cadiñanos, J., De Sandre-Giovannoli, A., Bernard, R.,
Courrier, S., Boccaccio, I., Boyer, A., Kleijer, WJ., Wagner, A., Giuliano, F.
et al. (2005) Loss of ZMPSTE24 (FACE-1) causes autosomal recessive
restrictive dermopathy and accumulation of lamin A precursors. Hum.
Mol. Genet. 14, 1503–1513
12 Novelli, G., Muchir, A., Sangiuolo, F., Helbling-Leclerc, A., D’Apice, M.R.,
Massart, C., Capon, F., Sbraccia, P., Federici, M., Lauro, R. et al. (2002)
Mandibuloacral dysplasia is caused by a mutation in LMNA-encoding
lamin A/C. Am. J. Hum. Genet. 71, 426–431
13 Puente, X.S., Quesada, V., Osorio, F.G., Cabanillas, R., Cadiñanos, J., Fraile,
J.M., Ordóñez, G.R., Puente, D.A., Gutiérrez-Fernández, A.,
Fanjul-Fernández, M. et al. (2011) Exome sequencing and functional
analysis identifies BANF1 mutation as the cause of a hereditary
progeroid syndrome. Am. J. Hum. Genet. 88, 650–656
14 Varela, I., Pereira, S., Ugalde, A.P., Navarro, C.L., Suárez, M.F., Cau, P.,
Cadiñanos, J., Osorio, F.G., Foray, N., Cobo, J. et al. (2008) Combined
treatment with statins and aminobisphosphonates extends longevity in
a mouse model of human premature aging. Nat. Med. 14, 767–772
C The
C 2011 Biochemical Society
Authors Journal compilation 1717
1718
Biochemical Society Transactions (2011) Volume 39, part 6
15 Liu, B.H., Wang, J.M., Chan, K.M., Tjia, W.M., Deng, W., Guan, X.Y., Huang,
J.D., Li, K.M., Chau, P.Y., Chen, D.J. et al. (2005) Genomic instability in
laminopathy-based premature aging. Nat. Med. 11, 780–785
16 Manju, K., Muralikrishna, B. and Parniak, V.K. (2006) Expression of
disease-causing lamin A mutants impairs the formation of DNA repair
foci. J. Cell Sci. 119, 2704–2714
17 diMasi, A., D’Apice, M.R., Ricordy, R., Tanzarella, C. and Novelli, G (2008)
The R527H mutation in LMNA gene causes an increased sensitivity to
ionizing radiation. Cell Cycle 7, 2030–2037
18 Allsopp, R.C, Vaziri, H., Patterson, C., Goldstein, S., Younglai, E.V., Futcher,
A.B., Greider, C.W. and Harley, C.B. (1992) Telomere length predicts
replicative capacity of human fibroblasts. Proc. Natl. Acad. Sci. U.S.A. 89,
10114–10118
19 Huang, S., Risques, R.A., Martin, G.M., Rabinovitch, P.S. and Oshima, J.
(2008) Accelerated telomere shortening and replicative senescence in
human fibroblasts overexpressing mutant and wild-type lamin A. Exp.
Cell Res. 314, 82–91
20 Liu, Y.Y., Rusinol, A., Sinensky, M., Wang, Y.J. and Zou, Y. (2006) DNA
damage responses in progeroid syndromes arise from defective
maturation of prelamin A. J. Cell Sci. 119, 4644–4649
21 Yang, S.H., Andres, D.A., Spielmann, H.P., Young, S.G. and Fong, L.G.
(2008) Progerin elicits disease phenotypes of progeria in mice whether
or not it is farnesylated. J. Clin. Invest. 118, 3291–3300
22 Bar, D.Z. and Gruenbaum, Y. (2010) Reversal of age-dependent nuclear
morphology by inhibition of prenylation does not affect lifespan in
Caenorhabditis elegans. Nucleus 1, 499–505
23 Varela, I., Cadiñanos, J., Pendás, AM., Gutiérrez-Fernández, A., Folgueras,
A.R., Sánchez, L.M., Zhou, Z.J., Rodrı́guez, F.J., Stewart, C.L., Vega, J.A.
et al. (2005) Accelerated ageing in mice deficient in Zmpste24 protease
is linked to p53 signalling activation. Nature 437, 564–568
24 Chowdhury, D., Dipanjan, M.C., Ishii, H., Peterson, C.L., Buratowski, S. and
Lieberman, J. (2005) γ -H2AX dephosphorylation by protein phosphatase
2A facilitates DNA double-strand break repair. Mol. Cell 20, 801–809
25 Pei, H., Zhang, L., Luo, K., Qin, Y., Chesi, M., Fei, F., Bergsagel, P.L., Wang,
L., You, Z. and Lou, Z. (2011) MMSET regulates histone H4K20
methylation and 53BP1 accumulation at DNA damage sites. Nature 470,
124–128
26 Tatum, D. and Li, S. (2011) Evidence that the histone methyltransferase
Dot1 mediates global genomic repair by methylating histone H3 on
lysine 79. J. Biol. Chem. 286, 17530–17535
27 Scaffidi, P. and Misteli, T. (2006) Lamin A-dependent nuclear defects in
human aging. Science 312, 1059–1063
28 Shumaker, D.K., Dechat, T., Kohlmaier, A., Adam, S.A., Bozovsky, M.R.,
Erdos, M.R., Eriksson, M., Goldman, A.E., Khuon, S., Collins, F.S. et al.
(2006) Mutant nuclear lamin A leads to progressive alterations of
epigenetic control in premature aging Proc. Natl. Acad. Sci. U.S.A. 103,
8703–8708
29 Zhang, Y., Ng, H.H., Erdjument-Bromage, H., Tempst, P., Bird, A. and
Reinberg, D. (1999) Analysis of the NuRD subunits reveals a histone
deacetylase core complex and a connection with DNA methylation.
Genes Dev. 13, 1924–1935
C The
C 2011 Biochemical Society
Authors Journal compilation 30 Pegoraro, G., Kubben, N., Wickert, U., Gohler, H., Hoffman, K. and
Mistelli, T. (2009) Ageing-related chromatin defects through loss of the
NURD complex. Nat. Cell Biol. 11, 1261–1267
31 Krishnan, V., Zi, M., Chow, Y., Wang, Z., Zhang, L., Liu, B., Liu, X. and Zhou,
Z. (2011) Histone H4 lysine 16 hypoacetylation is associated with
defective DNA repair and premature senescence in Zmpste24-deficient
mice. Proc. Natl. Acad. Sci. U.S.A. 108, 12325–12330
32 Liu, Y., Wang, Y., Rusinol, A.E., Sinensky, MS., Liu, J., Shell, S.M. and Zou,
Y. (2008) Involvement of xeroderma pigmentosum group A (XPA) in
progeria arising from defective maturation of prelamin A. FASEB J. 22,
603–611
33 Richards, S.A., Muter, J., Ritchie, P., Lattanzi, G. and Hutchison, C.J. (2011)
The accumulation of un-repairable DNA damage in laminopathy
progeria fibroblasts is caused by ROS generation and is prevented by
treatment with N-acetyl cysteine. Hum. Mol. Genet. 20, 3997–4004
34 Gonzalez-Suarez, I., Redwood, A.B., Perkins, S.M., Vermolen, B.,
Lichtensztejin, D., Grotsky, D., Morgado-Palacin, L., Gapud, E.J.,
Sleckman, B.P., Sullivan, T. et al. (2009) Novel roles for A-type lamins in
telomere biology and DNA damage response pathway. EMBO J. 28,
2414–2427
35 Dimitrova, N., Chen, Y.C., Spector, D.L. and de Lange, T. (2008) 53BP1
promotes non-homologous end joining of telomeres by increasing
chromatin mobility. Nature 456, 524–528
36 Verstraeten, V.L., Caputo, S., van Steensel, M.A., Duband-Goulet, I.,
Zinn-Justin, S., Kamps, M., Kuijpers, H.J., Ostlund, C., Worman, H.J., Briedé,
J.J. et al. (2009) The R439C mutation in LMNA causes lamin
oligomerization and susceptibility to oxidative stress. J. Cell. Mol. Med.
13, 959–971
37 Viteri, G., Chung, Y.W. and Stadtman, E.R. (2010) Effect of progerin on
the accumulation of oxidized proteins in fibroblasts from
Hutchinson–Gilford progeria patients. Mech. Ageing Dev. 131, 2–8
38 Lee, S.F. and Pervaiz, S. (2011) Assessment of oxidative stress-induced
DNA damage by immunofluorescence analysis of 8-oxodG. Methods Cell
Biol. 103, 99–113
39 Von Zglinicki, T., Saretzki, G., Ladhoff, J., Di Fagagna, F.D. and Jackson, S.P.
(2005) Human cell senescence as a DNA damage response. Mech.
Ageing Dev. 126, 111–117
40 Sedelnikova, O.A., Horikawa, I., Zimonjic, D.B., Popescu, N.C., Bonner,
W.M. and Barrett, J.C. (2004) Senescing human cells and ageing mice
accumulate DNA lesions with unrepairable double-strand breaks. Nat.
Cell Biol. 6, 168–170
41 Ukekawa, R., Miki, K., Fujii, M., Hirano, H. and Ayusawa, D. (2007)
Accumulation of multiple forms of lamin A with down-regulation of
FACE-1 suppresses growth in senescent human cells. Genes Cells 12,
397–406
Received 5 August 2011
doi:10.1042/BST20110700