Reprint of “DNA, the central molecule of aging”

Reprint of “DNA, the central molecule of aging”

G Model ARTICLE IN PRESS MUT-11549; No. of Pages 7 Mutation Research xxx (2016) xxx–xxx Contents lists available at ScienceDirect Mutation Resear...

1MB Sizes 1 Downloads 69 Views

G Model

ARTICLE IN PRESS

MUT-11549; No. of Pages 7

Mutation Research xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis journal homepage: www.elsevier.com/locate/molmut Community address: www.elsevier.com/locate/mutres

Review

Reprint of “DNA, the central molecule of aging”夽 Peter Lenart a , Lumir Krejci a,b,c,∗ a

Department of Biology, Masaryk University, Brno, Czech Republic International Clinical Research Center, Center for Biomolecular and Cellular Engineering, St. Anne’s University Hospital Brno, Brno, Czech Republic c National Centre for Biomolecular Research, Masaryk University, Brno, Czech Republic b

a r t i c l e

i n f o

Article history: Received 17 December 2015 Received in revised form 16 January 2016 Accepted 30 January 2016 Available online xxx Keywords: DNA Aging Chromatin structure Telomeres DNA damage DNA repair Mutagenesis

a b s t r a c t Understanding the molecular mechanism of aging could have enormous medical implications. Despite a century of research, however, there is no universally accepted theory regarding the molecular basis of aging. On the other hand, there is plentiful evidence suggesting that DNA constitutes the central molecule in this process. Here, we review the roles of chromatin structure, DNA damage, and shortening of telomeres in aging and propose a hypothesis for how their interplay leads to aging phenotypes. © 2016 Published by Elsevier B.V.

Contents 1.

2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 1.1. Chromatin structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 1.2. Telomeres shortening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 1.3. DNA damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

1. Introduction Aging is a complex biological process resulting in the decline of almost all physiological functions, which in turn leads to a time-dependent increase in mortality. Many theories have tried to explain the aging process, but none is universally accepted [1,2]. Although many biomolecules could play a role in aging,

DOI of original article: http://dx.doi.org/10.1016/j.mrfmmm.2016.01.007. 夽 This article is a reprint of a previously published article. For citation purposes, please use the original publication details “Mutation Research 786 (2016) 1–7”. ∗ Corresponding author at: Department of Biology, Faculty of Medicine, Masaryk University, Kamenice 5/A7, Brno 62500, Czech Republic. Fax: +420-549-492-556. E-mail address: [email protected] (L. Krejci).

DNA seems to be the most relevant. This is supported by several arguments. First, syndromes of accelerated aging are often associated with defects in DNA repair genes [3,4]. Second, changes of chromatin structure, shortening of telomeres and accumulation of DNA damage are all associated with aging and life span [5–9]. These chromosomal changes do not act in isolation but are rather tightly interconnected. Changes of chromatin structure can accelerate shortening of telomeres [10] alter susceptibility to DNA damage [11] and modify transcription [12] thus influencing almost all cellular functions. Vice versa, DNA damage can lead to changes of chromatin structure [13] and accelerate telomere shortening [14]. Here we review how these three types of chromosomal changes and their interplay influence the aging process.

http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002 0027-5107/© 2016 Published by Elsevier B.V.

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

2

1.1. Chromatin structure Chromatin is a nucleoprotein complex that can be understood as a dynamic, three-dimensional, higher-order structural state of a chromosome. The basic unit of chromatin is a nucleosome consisting of four pairs of core histones (H2A, H2B, H3, H4) around which 147 bases of DNA are wrapped [15]. Higher-order chromatin structure is also formed by linker H1 histone and other non-histone proteins. There exists two basic types of chromatin: highly condensed heterochromatin or loosly condensed euchromatin, characterized by typically transcriptional inactivity and resistance to DNA damage or transcriptional activity and susceptibility to DNA damage, respectively [11]. Therefore, it is not surprising that several studies have shown depletion of core histones to be associated with aging. In yeast, histone concentration decreases with age and seems to be directly related to aging, since their overexpression leads to a 65 % increase in replicative life span [8]. The possible mechanism might be changes in transcription, as loss of nucleosome in yeast was reported to cause globally increased gene expression [16]. In addition, decreases in H3 and H4 histone levels also have been observed in human senescent fibroblasts, where their concentrations were reported as reduced by half compared to levels in young cells [5,17]. Although synthesis of new histones diminishes with age, it is notable that their expression actually increases [18]. Even as cells are trying to stabilize histone levels, therefore, this might be counteracted by an increased rate of mRNA degradation or increasingly ineffective translation. It would therefore be interesting to test whether overexpression of histones would extend the life span of human cells. Chromatin structure can be altered not only by changes in the number of histones but also by their post-translational modifications. More than 60 modification sites have been identified on histones, which, together with the 8 basic types of modifications, means the number of specific modifications (type and position) is immense. Histone modifications are known to play important roles in regulation of many cellular processes such as replication, transcription, and DNA repair [19], and so it is not surprising that types and amounts of modifications change with age. For instance, acetylation of lysine 16 on histone 4 (H4K16ac) increases with age in yeast [5,20]. Other examples include increases in H3K9ac and reductions in H3K56ac [5]. Acetylation in general removes positive charge from lysine and could be expected to weaken the binding of DNA to histone and, in turn, make chromatin structure looser. Indeed, H4K16ac has been implicated in determining chromatin structure and influencing the interaction between nonhistone proteins and chromatin [21,22]. In addition, the dynamic status of the modification plays an important role, as Sir2, the main deacetylase regulating this modification, is also known to extend life span in several invertebrates such as Saccharomyces cerevisiae [23], Caenorhabditis elegans [24], and Drosophila melanogaster [25]. It is not clear, however, if SIRT1, the human Sir2 orthologue [26,27], can extend life span in mammals, since overexpressing SIRT1 in all mouse tissues was shown to have no life-extending effects despite its having a positive effect on several pathologies associated with aging [28]. Surprisingly, another study overexpressing SIRT1 in mouse brains reported 11 % elongation of life span [29]. It can therefore be speculated that SIRT1 is beneficial to life span only in certain tissues or only at low concentrations. Strong evidence for a connection of histone acetylation with chromatin maintenance and aging can be seen in the effect of the polyamine known as spermidine on life span and aging. In general, polyamines are associated with cell growth [30], their depletion inhibits apoptosis [31], and they are also implicated in carcinogenesis inasmuch as their concentrations are elevated in cancer cells [32]. In addition, it has been shown that yeast as well as

mammalian cells synthesize less polyamines with age. Eisenberg et al. have shown that spermidine supplementation extends life span in yeast, nematodes, flies, mice, and also human cells and that this effect is accompanied by hypoacetylation of H3K9, H3K14 and H3K18. Accordingly, inactivation of acetyltransferases responsible for acetylation of these lysines also extended the life span of yeast and decreased the effect of spermidine treatment [33]. Histones modification by methylation of their lysine or arginine residues plays also important role. In contrast to the histone acetylation, methylation can be associated with either active or repressed transcription, depending on the affected residue [34]. While trimethylation of lysine 4 on histone 3 is associated with active transcription [35], trimethlyation of lysine 27 on the same histone is associated with repressed transcription [34]. Both of these modifications are linked to aging. In human senescent fibroblasts H3K4me3 is enriched and occupies new parts of genome [36]. Spreading of H3K4me3 during aging has bean also observed in mouse hematopoietic stem cells [37]. The most direct evidence of role of H3K4me3 in aging comes from C. elegans, where overexpression of RBR-2, the H3K4me3 demethylase, increase life span, while its knockdown has opposite effect. Furthermore knockdown or mutation of genes encoding H3K4 methyltransferases increases life span [38]. In contrast to H3K4me3 mark, H3K27me3 decreases during aging and knockdown of UTX-1, H3K27me3 demethylase extends life span of C. elegans [39]. However, role of H3K27me3 in aging is not as clear, as experiments in D. melanogaster have shown that mutation in H3K27-specific methyltransferase E(Z) increases life span of flies and reduce amount of H3K27me3 [40]. Other histone modifications are less characterized, but may nevertheless influence aging. For example low levels of ubiquitination of H2B were found to be necessary for yeast cells to attain normal life span possibly trough regulating Sir2 recruitment [41]. DNA methylation is also relevant in determining chromatin structure [42,43] and regulating gene expression [44]. Changes in methylation are typical for cancer cells, but similar changes have been observed also in senescent cells [45–47]. Aging cells exhibit global hypomethylation and local hypermethylation. While hypomethylation is typical for noncoding parts of the genome, sequences near the promotors of several genes regulating the cell cycle are often hypermethylated [48]. Several authors have been able recently to use methylation patterns to successfully predict the age of several different tissues, mostly by analyzing methylation of different CpG sites as an aging clock [49–51]. Furthermore, the methylation patterns of genomic DNA were also shown to predict mortality regardless of current health status, life style, and known genetic predispositions [52], thus suggesting a direct relationship between DNA methylation and aging. A causal role of DNA methylation in aging is also suggested by the ability of methionine restriction – a well-known dietary intervention – to prolong the life span of various model organisms [53–56] and human fibroblast [57]. It is expected that this is due to methionine’s serving as a substrate for methyl-transferases and thus affecting the methylation state [58]. 1.2. Telomeres shortening Telomeres are nucleoprotein complexes protecting coding sequences from replicative shortening of chromosomes. Telomeric DNA consists of a G-rich repetitive sequence (TTAGGG in mammals) bound by numerous proteins forming a shelterin complex which also blocks these ends from being recognized as DNA double-strand breaks (DSBs) [59]. Telomeres progressively shorten after each cell division, limiting the number of divisions of somatic cells. They can be extended by a special enzyme termed telomerase, which in humans is active mostly in embryonic stem cells [60]. Telomeres and telomerase have long been implicated in the aging process,

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

as introduction of telomerase to cells in vitro is able to postpone senescence and promote immortalization [6]. It was later proven that transient telomerase expression did extend the life span of human cells, but this did not result in their immortalization [61]. This led to the hypothesis that telomere shortening is the primary cause of aging and sparked many studies analyzing the impact of telomere length on human health. The first studies defining the role of telomeres in aging of mammals involved mice with knock-out of the genes necessary for telomerase synthesis (Terc or Tert genes). Surprisingly, firstgeneration mice from these knock-out lines are healthy with no significant changes in phenotype [62,63], thus indicating that telomerase is not necessary for life and that a lack of phenotype might be due to residual telomerase activity in the organism. Accordingly, the phenotype should be more visible in future generations. Indeed, from the third generation onward these mice have significantly shorter telomeres, decreased fertility, overall frailty, and atrophy of tissues leading to organ malfunctions [64]. Aging is exacerbated in mice defective for the WRN helicase, mutated in Werner syndrome and responsible for telomere maintenance [65,66]. This is interesting since Werner syndrome is a well-known progeria in humans showing an interaction with different genome maintenance mechanisms [67]. Telomerase is often expressed in cancer cells, since it can allow unrestricted proliferation, and it is therefore not surprising that classical transgenesis increases the incidence of cancer [68–70]. Nevertheless, overexpression of telomerase in mice genetically modified to be more resistant to cancer by increased expression of tumor-suppressor genes led to slower aging and increase in median life span by 40 % [71]. Additionally, cancer incidence is not increased if the telomerase is transduced into adult mice and transduction of telomerase into 1-year-old mice extended their life by 24 % [72]. While the aforementioned findings clearly demonstrate that telomere shortening and accordingly telomerase activity influence aging, the results of comparative studies show that telomere shortening is almost certainly not the primary cause of aging. For example, a study comparing telomerase activity in different rodents found no correlation between this enzyme’s activity and rodent life span. On the other hand, a significant negative correlation with body weight was observed and suggests a possible cancerprotective strategy of organisms [73]. An identical relationship between telomerase activity and body weight was also observed in a study comparing more than 60 mammalian species [74]. Interestingly, a negative correlation between telomere length and life span was found. On the other hand, telomere length can be used to predict the life span of individual zebra finches [75]. Although this seemingly contradictory data could be used to argue that telomeres are rather downstream effectors of aging, in species that evolved to use their proliferation-limiting potential as a protection mechanism against cancer their telomere shortening speed can determine individual life span. Studies in humans analyzing the relationship between telomere length and health and mortality have led to findings no less interesting. The initial study analyzing 143 subjects older than 60 showed a strong correlation between shorter telomeres in leukocytes from periphery blood and increased mortality [76]. This result was not confirmed, however, by more recent studies involving considerably larger cohorts. While a study on a cohort of 3,075 healthy men and women aged 70–79 found no significant correlation between telomere length and mortality, it did show a correlation between telomere length and years of healthy life span [77]. Short telomeres were also associated with risks of many illnesses such as various types of cancer [78], renal dysfunction [79], autism [80], mortality from cancer in patients with chronic obstructive pulmonary disease [81], and many others. In principle, however, this kind of study is only correlative and cannot prove a

3

causal relationship. The situation is also complicated by the fact that longer telomeres were also associated with risks of deadly diseases. While one study found a correlation between long telomeres and a risk of lung cancer in never-smoking women [82], another associated longer telomeres with susceptibility to colorectal carcinoma [83]. Telomere length greater than average was also correlated with a risk of breast cancer [84]. This brief overview illustrates that within the current state of knowledge it remains very difficult or even impossible to objectively evaluate the influence of telomere length on human health. Therefore, a large meta-analysis is needed that could provide more clear evidence. 1.3. DNA damage As a consequence of endogenous or exogenous influences, the chemical structure of DNA can be altered or lead to breakage of one or both DNA strands. A causal role of DNA damage in aging is strongly suggested by its role in progeria syndromes [3] as well as other indications. One of these is that many life-span-extending strategies lead to enhanced DNA repair or decreased DNA damage. For example, caloric restriction, the oldest known intervention able to prolong life span, has been reported to decrease age-related decline of base excision repair (BER) [85], nucleotide excision repair [86], and non-homologous end joining [87]. It also increases expression of SIRT1 [88,89], which enhances function of homologous recombination [90]. Furthermore the mTOR inhibitor, rapamycin was shown in mice to increase life span [91] and reduce DNA damage and cancer in skin exposed to a DNA damaging agent [92]. DNA adducts are a common type of DNA damage. Formation of DNA adducts is caused mainly by reactive-oxygen species, with 8-oxoguanine (8-oxoG) being commonly used as a marker of oxidative stress [93]. Current data suggest that DNA adducts do not play an important role in aging since even a 20-fold increase in 8-oxoG levels in mice did not result in visibly accelerated aging [94]. Another frequent type of DNA damage involves apurinic/apyrimidinic-sites (AP sites), which arise from spontaneous depurination or through repair of modified bases via BER. Creation of AP sites eventually leads to base replacements, frame shift mutations and DNA breaks, and accumulation of AP sites with age is supported by their 7-fold increase in leukocytes from old donors when compared to young donors [95]. In addition, BER is required for yeast to attain a full chronological life span [96], supporting its role in aging. Single-strand breaks (SSBs) are the most common type of DNA damage, with more than 10,000 lesions arising per cell per day [97]. This important type of DNA damage is probably not connected to aging, however, since studies analyzing SSBs in mice have found no age-dependent derivation in SSB levels [98,99]. DSBs are the most dangerous type of damage and arguably are also the most associated with aging. Unrepaired or incorrectly repaired DSBs can lead to loss of a part of a chromosome or to gross chromosomal reengagements and genomic instability, which processes are linked with many diseases including those related to advanced age and DSB repair defects can lead to early aging in mice [100]. DSBs increases with age both in vitro and in vivo [101,7], as does the time within which cells respond to DSBs, thus indicating the gradual loss of DNA repair efficiency [102]. Accordingly, it was recently shown that inducing DSBs in mice livers by expression of the restriction enzyme Sac1 accelerated this organ’s aging [9]. DNA damage response (DDR) is a mechanism by which cells sense damage and trigger corresponding repair. This response includes modification of histones near the damage site. Specifically, H2AX histones are phosphorylated (␥–H2AX) [103–105]. These changes can span as many as even several megabases from the DSBs [106] and as a consequence can also inhibit transcription [107,108]. However, heterochromatin, which is also induced by aging, limits DDR [109,110], and this can be attributed to an inability of cells

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7 4

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

Fig. 1. Interaction between different types of molecular damage and their relationship to aging. DNA damage arises as a consequence of intrinsic or extrinsic influences and when not repaired correctly can lead to mutations. Mutations alone probably do not play a causal role in aging, but they can lead to other biological changes and pathologies, including cancer development. On the other hand, the actual repair process leads to changes in chromatin structure, which can, in turn, make chromatin structure more prone to additional DNA damage. The main result of changes in chromatin structure, however, is deregulation of transcription, which may result in modification of tissue-specific expression patterns and thereby decrease the biological functions of tissues and lead to accelerated aging. Deregulated transcription can also impair DNA damage response machinery and chromatin structure maintenance, because specific expression profiles are needed for these processes to ensure maximum efficiency. Most importantly, it is not the effect of these individual processes alone, but rather their combination that might result in the aging phenotype.

to recognize DNA damage. Such assumption has been challenged, however, since ␥–H2AX can be generated successfully even in the very dense structure of mitotic chromosomes [111]. Another possible explanation might be that this reflects changes in chromatin structure rather than recognition of DNA damage itself by DNA repair proteins [112,113]. A variety of other chromatin modifications also accumulate at sites of DNA damage. Very important is ubiquitination of H2A and H2B that further stimulates cell response to DNA damage [114,115]. These modifications not only play a role in DNA repair but also influence transcription both locally and globally, and many changes in transcription caused by DNA damage are transitional [116]. In some cases, however, inability to restore a normal chromatin structure after DNA repair can significantly deregulate transcription. One such example is delocalization of the histone deacetylase SIRT1. In response to DNA damage, SIRT1 binds to the damaged site. There, it plays a role in repair, although this leads to its deficiency in other parts of the genome and the depression of SIRT1 regulated genes [13]. This process is similar to changes observed in aging and it seems to be evolutionarily conserved inasmuch as binding of SIR proteins to DNA damage sites also leads to their depression in other parts of the genome and manifestation of an aging phenotype [117]. Other proteins, too, are implicated in DNA repair and chromatin structure that also regulate life span. Both SIRT1 and yeast Sir2 have important roles in DDR, for example, as they bind to the DNA damage site and there deacetylate several proteins involved in DNA repair and stimulate their activity [118]. DSB repair leads to acetylation of the N-terminal lysines of histones H3 and H4 [117]. This is very intriguing with regard to aging because, as mentioned above, hypoacetylation of the N-terminal lysines of histones is at least partially responsible for the life-extending effects of spermidine [33]. Sirtuins also play an important role in maintaining heterochromatin on repetitive sequences of rDNA and telomeres and prevent aberrant recombination between these repetitive sequences [90,119]. As would be expected, mice with inactive SIRT6 have increased genomic instability and show accelerated aging [120]. In addition, overexpression of pch-2, which modulates meiotic recombination and helps to maintain ribosomal DNA stability [121] was recently shown to extend the life span of C. elegans and increase resistance to multiple stressors, thereby affecting

both DNA and protein integrity. Furthermore, even though the effect of pch-2 modulation is sir-2 independent, the same authors propose a mechanism of pch-2 and sir-2 interaction leading to life extension [122]. In addition, deletion of Exo1, nuclease involved in mismatch repair and recombination, is able to prolong the life span of telomere-dysfunctional mice, possibly by impairing DNA damage signals at DNA breaks [123]. Mutations arise as a consequence of nonfunctional DNA repair or mistakes during replication. Mutations were first mentioned as the primary cause of aging in the middle of the 20th century [124,125]. It was nevertheless impossible for decades to directly compare mutation rate in differently aged cells. Finally, several studies showed that mutations are indeed accumulating with age and reflect the proliferative capacity of tissues [126–128]. Despite this, small mutations that result from improper repair of base lesions, base mismatches or indels are unlikely to constitute one of the primary causes of aging since transgenic mice with defects in mismatch repair [129] or antioxidant defense [130] exhibit an increase in spontaneous mutation rate, but do not shown any signs of accelerated aging apart from increased cancer incidence. Similarly mice, with defective transcription-coupled repair exhibited early ageing without an increase in mutation frequency [131]. A recent study also confirmed that aging in S. cerevisiae is not caused by an increase in small mutations, since old yeast cells with few or no mutations were dying even though young cells mutated in eight thiol peroxidases, which are enzymes detoxifying hydroperoxides, had twice as many mutations and continued dividing [132]. By contrast, defective repair of DNA DSBs could accelerate aging since mutations in multiple DSB repair genes leads to early aging in mice and people and since chromosomal abnormalities correlate with age [100]. Thus, the accumulation of chromosomal rearrangements might be a good indicator for some age-associated pathologies, such as cancer. 2. Conclusion It is unlikely for such a complex process as aging to be driven just by one specific kind of disturbance. Accordingly, we suggest that aging is caused by the interplay of several types of damage. Even though their relationship may form complex network which is yet

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

to be fully understood, we believe that there is already a traceable pattern suggesting a possible order of their interaction (Fig. 1). We propose that DNA damage serves as trigger of aging. Accumulation of DNA damage, mainly DSBs, leads to chromatin modifications and chromatin’s loosening, which in turn makes chromatin more prone to additional DNA damage. These changes of chromatin structure then affect transcription, and accelerate telomere shortening. Faster telomere shortening may then lead to quicker depletion of stem cells in organisms thus reducing regenerative capacity of tissues. In addition, we suggest that changes in transcription can lead to age-related decrease in physiological functions, since cells with progressively more deregulated transcription can scarcely maintain optimal biological function. Moreover, this notion is strongly supported by fact that changing levels of specific histone modification also influences life span. Finally, unrepaired or incorrectly repaired DNA damage can also lead to mutations, and these can inversely affect all the aforementioned processes. Even though accumulation of small mutations probably does not play a causal role in aging, it can influence any of the individual processes and thus manifest some age-related pathologies, and especially increased incidence of cancer. The accumulation of chromosomal rearrangements could impact aging by directly impairing cell homeostasis or inducing cell senescence. Most importantly, this model indicates not only the possible importance of these processes individually but in particular their combination that eventually leads to aging. It might be this combination that gives to the entire process the complexity that would lead to a determination of direct causality. Nevertheless, it can be expected that future research in this area will shed more light on how changes relating to the DNA molecule affect the aging process. A better understanding of the mechanism will enable us to identify appropriate targets for intervention, generate new compounds, and develop new strategies to counteract aging phenotypes. Furthermore, it could be expected that aging research will continue to capture the attention of an ever-increasing number of scientists, since mathematical models clearly demonstrate that to delay aging will have enormous health and economic benefits greatly outweighing the potential benefits of separately addressing heart disease or cancer [133]. Conflict of interest The authors declare no conflict of interest. Acknowledgements This work was supported by Czech Science Foundation grants GACR13-26629S and GACR207/12/2323, European Regional Development Fund [Project FNUSA-ICRC, no. CZ.1.05/1.1.00/02.0123] and Research Support Programme [GAMU] – MUNI/M/1894/2014. References [1] K. Jin, Modern biological theories of aging, Aging Dis. 1 (2010) 72–74. [2] L.S. Trindade, T. Aigaki, A.A. Peixoto, A. Balduino, I.B. Mânica da Cruz, J.G. Heddle, A novel classification system for evolutionary aging theories, Front Genet. (2013) 4. [3] C.R. Burtner, B.K. Kennedy, Progeria syndromes and ageing: what is the connection? Nat. Rev. Mol. Cell Biol. 11 (2010) 567–578. [4] A.A. Freitas, O. Vasieva, Magalhães J.P de, A data mining approach for classifying DNA repair genes into ageing-related or non-ageing-related, BMC Genomics 12 (2011) 27. [5] R.J. O’sullivan, S. Kubicek, S.L. Schreiber, J. Karlseder, Reduced histone biosynthesis and chromatin changes arising from a damage signal at telomeres, Nat. Struct. Mol. Biol. 17 (2010) 1218–1225. [6] A.G. Bodnar, M. Ouellette, M. Frolkis, S.E. Holt, C.P. Chiu, G.B. Morin, et al., Extension of life-span by introduction of telomerase into normal human cells, Science 279 (1998) 349–352.

5

[7] O.A. Sedelnikova, I. Horikawa, D.B. Zimonjic, N.C. Popescu, W.M. Bonner, J.C. Barrett, Senescing human cells and ageing mice accumulate DNA lesions with unrepairable double-strand breaks, Nat. Cell Biol. 6 (2004) 168–170. [8] J. Feser, D. Truong, C. Das, J.J. Carson, J. Kieft, T. Harkness, et al., Elevated histone expression promotes life span extension, Mol. Cell 39 (2010) 724–735. [9] R.R. White, B. Milholland, A. de Bruin, S. Curran, R.-M. Laberge, H. van Steeg, et al., Controlled induction of DNA double-strand breaks in the mouse liver induces features of tissue ageing, Nat. Commun. (2015) 6. [10] I. Mozgová, P. Mokroˇs, J. Fajkus, Dysfunction of chromatin assembly factor 1 induces shortening of telomeres and loss of 45S rDNA in Arabidopsis thaliana[W][OA], Plant Cell 22 (2010) 2768–2780. [11] M. Falk, E. Lukáˇsová, S. Kozubek, Chromatin structure influences the sensitivity of DNA to -radiation, Biochim. Biophys. Acta BBA: Mol. Cell Res. 1783 (2008) 2398–2414. [12] B. Li, M. Carey, J.L. Workman, The role of chromatin during transcription, Cell 128 (2007) 707–719. [13] P. Oberdoerffer, S. Michan, M. McVay, R. Mostoslavsky, J. Vann, S.-K. Park, et al., DNA damage-induced alterations in chromatin contribute to genomic integrity and age-related changes in gene expression, Cell 135 (2008) 907–918. [14] S. Oikawa, S. Kawanishi, Site-specific DNA damage at GGG sequence by oxidative stress may accelerate telomere shortening, FEBS Lett. 453 (1999) 365–368. [15] K. Luger, A.W. Mäder, R.K. Richmond, D.F. Sargent, T.J. Richmond, Crystal structure of the nucleosome core particle at 2.8 Å resolution, Nature 389 (1997) 251–260. [16] Z. Hu, K. Chen, Z. Xia, M. Chavez, S. Pal, J.-H. Seol, et al., Nucleosome loss leads to global transcriptional up-regulation and genomic instability during yeast aging: Genes Dev 2014;28:396–408. telomeres, Nat. Struct. Mol. Biol. 17 (2010) 1218–1225. [17] A. Ivanov, J. Pawlikowski, I. Manoharan, J. van Tuyn, D.M. Nelson, T.S. Rai, et al., Lysosome-mediated processing of chromatin in senescence, J. Cell Biol. 202 (2013) 129–143. [18] I. Lesur, J.L. Campbell, The transcriptome of prematurely aging yeast cells is similar to that of telomerase-deficient cells, Mol. Biol. Cell 15 (2004) 1297–1312. [19] T. Kouzarides, Chromatin modifications and their function, Cell 128 (2007) 693–705. [20] W. Dang, K.K. Steffen, R. Perry, J.A. Dorsey, F.B. Johnson, A. Shilatifard, et al., Histone H4 lysine-16 acetylation regulates cellular lifespan, Nature 459 (2009) 802–807. [21] M. Shogren-Knaak, H. Ishii, J.-M. Sun, M.J. Pazin, J.R. Davie, C.L. Peterson, Histone H4-K16 acetylation controls chromatin structure and protein interactions, Science 311 (2006) 844–847. [22] B.-R. Zhou, H. Feng, R. Ghirlando, H. Kato, J. Gruschus, Y. Bai, Histone H4 K16Q mutation, an acetylation mimic, causes structural disorder of its N-terminal basic patch in the nucleosome, J. Mol. Biol. 421 (2012) 30–37. [23] D.A. Sinclair, L. Guarente, Extrachromosomal rDNA circles— a cause of aging in yeast, Cell 91 (1997) 1033–1042. [24] H.A. Tissenbaum, L. Guarente, Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans, Nature 410 (2001) 227–230. [25] B. Rogina, S.L. Helfand, Sir2 mediates longevity in the fly through a pathway related to calorie restriction, Proc. Natl.Acad. Sci. U. S. A 101 (2004) 15998–16003. [26] A. Vaquero, M. Scher, D. Lee, H. Erdjument-Bromage, P. Tempst, D. Reinberg, Human SirT1 interacts with histone H1 and promotes formation of facultative heterochromatin, Mol. Cell 16 (2004) 93–105. [27] K. Pruitt, R.L. Zinn, J.E. Ohm, K.M. McGarvey, S.-H.L. Kang, D.N. Watkins, et al., Inhibition of SIRT1 reactivates silenced cancer genes without loss of promoter DNA hypermethylation, PLoS Genet. (2006) 2. ˜ ˜ M. Canamero, F. Mulero, B. Martinez-Pastor, O. [28] D. Herranz, M. Munoz-Martin, Fernandez-Capetillo, et al., Sirt1 improves healthy ageing and protects from metabolic syndrome-associated cancer, Nat. Commun. 1 (2010) 3. [29] A. Satoh, C.S. Brace, N. Rensing, P. Clifton, D.F. Wozniak, E.D. Herzog, et al., Sirt1 extends life span and delays aging in mice through the regulation of Nk2 homeobox 1 in the DMH and LH, Cell Metab. 18 (2013) 416–430. [30] K. Igarashi, K. Kashiwagi, Modulation of cellular function by polyamines, Int. J. Biochem. Cell Biol. 42 (2010) 39–51. [31] F. Flamigni, I. Stanic’, A. Facchini, S. Cetrullo, B. Tantini, R.M. Borzì, et al., Polyamine biosynthesis as a target to inhibit apoptosis of non-tumoral cells, Amino Acids 33 (2007) 197–202. [32] E.W. Gerner, F.L. Meyskens, Polyamines and cancer: old molecules, new understanding, Nat. Rev. Cancer 4 (2004) 781–792. [33] T. Eisenberg, H. Knauer, A. Schauer, S. Büttner, C. Ruckenstuhl, D. CarmonaGutierrez, et al., Induction of autophagy by spermidine promotes longevity, Nat. Cell Biol. 11 (2009) 1305–1314. [34] E.L. Greer, Y. Shi, Histone methylation: a dynamic mark in health, disease and inheritance, Nat. Rev. Genet. 13 (2012) 343–357. [35] H. Santos-Rosa, R. Schneider, A.J. Bannister, J. Sherriff, B.E. Bernstein, N.C.T. Emre, et al., Active genes are tri-methylated at K4 of histone H3, Nature 419 (2002) 407–411. [36] P.P. Shah, G. Donahue, G.L. Otte, B.C. Capell, D.M. Nelson, K. Cao, et al., Lamin B1 depletion in senescent cells triggers large-scale changes in gene expression and the chromatin landscape, Genes Dev. 27 (2013) 1787–1799.

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7 6

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

[37] D. Sun, M. Luo, M. Jeong, B. Rodriguez, Z. Xia, R. Hannah, et al., Epigenomic profiling of young and aged HSCs reveals concerted changes during aging that reinforce self-renewal, Cell Stem Cell 14 (2014) 673–688. [38] E.L. Greer, T.J. Maures, A.G. Hauswirth, E.M. Green, D.S. Leeman, G.S. Maro, et al., Members of the histone H3 lysine 4 trimethylation complex regulate lifespan in a germline-dependent manner in C. elegans, Nature 466 (2010) 383–387. [39] T.J. Maures, E.L. Greer, A.G. Hauswirth, A. Brunet, H3K27 demethylase UTX1 regulates C. elegans lifespan in a germline indipedent, insulin-dependent, manner, Aging Cell 10 (2011) 980–990. [40] A.P. Siebold, R. Banerjee, F. Tie, D.L. Kiss, J. Moskowitz, P.J. Harte, Polycomb repressive complex 2 and trithorax modulate drosophila longevity and stress resistance, Proc. Natl. Acad. Sci. U. S. A. 107 (2010) 169–174. [41] B.H. Rhie, Y. Song, H. Ryu, S.H. Ahn, Cellular aging is associated with increased ubiquitylation of histone H2B in yeast telomeric heterochromatin, Biochem. Biophys. Res. Commun. 439 (2013) 570–575. [42] T. Hashimshony, J. Zhang, I. Keshet, M. Bustin, H. Cedar, The role of DNA methylation in setting up chromatin structure during development, Nat. Genet. 34 (2003) 187–192. [43] N. Gilbert, I. Thomson, S. Boyle, J. Allan, B. Ramsahoye, W.A. Bickmore, DNA methylation affects nuclear organization, histone modifications, and linker histone binding but not chromatin compaction, J. Cell Biol. 177 (2007) 401–411. [44] A.M. Deaton, A. Bird, CpG islands and the regulation of transcription, Genes Dev. 25 (2011) 1010–1022. [45] A.H. Ting, K.M. McGarvey, S.B. Baylin, The cancer epigenome—components and functional correlates, Genes Dev. 20 (2006) 3215–3231. [46] D. Sproul, R.R. Meehan, Genomic insights into cancer-associated aberrant CpG island hypermethylation, Brief. Funct. Genomics 12 (2013) 174–190. [47] G.C. Hon, R.D. Hawkins, O.L. Caballero, C. Lo, R. Lister, M. Pelizzola, et al., Global DNA hypomethylation coupled to repressive chromatin domain formation and gene silencing in breast cancer, Genome Res. 22 (2012) 246–258. [48] H.A. Cruickshanks, T. McBryan, D.M. Nelson, N.D. VanderKraats, P.P. Shah, J. van Tuyn, et al., Senescent cells harbour features of the cancer epigenome, Nat. Cell Biol. 15 (2013) 1495–1506. [49] S. Horvath, DNA methylation age of human tissues and cell types, Genome Biol. 14 (2013) R115. [50] G. Hannum, J. Guinney, L. Zhao, L. Zhang, G. Hughes, S. Sadda, et al., Genomewide methylation profiles reveal quantitative views of human aging rates, Mol. Cell 49 (2013) 359–367. [51] C.I. Weidner, Q. Lin, C.M. Koch, L. Eisele, F. Beier, P. Ziegler, et al., Aging of blood can be tracked by DNA methylation changes at just three CpG sites, Genome Biol. 15 (2014) R24. [52] R.E. Marioni, S. Shah, A.F. McRae, B.H. Chen, E. Colicino, S.E. Harris, et al., DNA methylation age of blood predicts all-cause mortality in later life, Genome Biol. (2015) 16. [53] B.C. Lee, A. Kaya, S. Ma, G. Kim, M.V. Gerashchenko, S.H. Yim, et al., Methionine restriction extends lifespan of Drosophila melanogaster under conditions of low amino acid status, Nat. Commun. 5 (2014) 3592. [54] J.A. Zimmerman, V. Malloy, R. Krajcik, N. Orentreich, Nutritional control of aging, Exp. Gerontol. 38 (2003) 47–52. [55] R.A. Miller, G. Buehner, Y. Chang, J.M. Harper, R. Sigler, M. Smith-Wheelock, Methionine-deficient diet extends mouse lifespan, slows immune and lens aging, alters glucose, T4, IGF-I and insulin levels, and increases hepatocyte MIF levels and stress resistance, Aging Cell 4 (2005) 119–125. [56] L. Sun, A.A. Sadighi Akha, R.A. Miller, J.M. Harper, Life-span extension in mice by preweaning food restriction and by methionine restriction in middle age, J. Gerontol. A: Biol. Sci. Med. Sci. 64A (2009) 711–722. [57] R. Kozieł, C. Ruckenstuhl, E. Albertini, M. Neuhaus, C. Netzberger, M. Bust, et al., Methionine restriction slows down senescence in human diploid fibroblasts, Aging Cell 13 (2014) 1038–1048. [58] I. Sanchez-Roman, A. Gomez, J. Gomez, H. Suarez, C. Sanchez, A. Naudi, et al., Forty percent methionine restriction lowers DNA methylation, complex I ROS generation, and oxidative damage to mtDNA and mitochondrial proteins in rat heart, J. Bioenerg. Biomembr. 43 (2011) 699–708. [59] Lange T. de, Shelterin: the protein complex that shapes and safeguards human telomeres, Genes Dev. 19 (2005) 2100–2110. [60] E. Hiyama, K. Hiyama, Telomere and telomerase in stem cells, Br. J. Cancer 96 (2007) 1020–1024. [61] S. Steinert, J.W. Shay, W.E. Wright, Transient expression of human telomerase extends the life span of normal human fibroblasts, Biochem. Biophys. Res. Commun. 273 (2000) 1095–1098. [62] M.A. Blasco, H.-W. Lee, M.P. Hande, E. Samper, P.M. Lansdorp, R.A. DePinho, et al., Telomere shortening and tumor formation by mouse cells lacking telomerase RNA, Cell 91 (1997) 25–34. [63] P.A. Farazi, J. Glickman, J. Horner, R.A. DePinho, Cooperative interactions of p53 mutation, telomere dysfunction, and chronic liver damage in hepatocellular carcinoma progression, Cancer Res. 66 (2006) 4766–4773. [64] H.-W. Lee, M.A. Blasco, G.J. Gottlieb, J.W. Horner, C.W. Greider, R.A. DePinho, Essential role of mouse telomerase in highly proliferative organs, Nature 392 (1998) 569–574. [65] X. Du, J. Shen, N. Kugan, E.E. Furth, D.B. Lombard, C. Cheung, et al., Telomere shortening exposes functions for the mouse Werner and Bloom syndrome genes, Mol. Cell Biol. 24 (2004) 8437–8446.

[66] S. Change, A.S. Multani, N.G. Cabrera, M.L. Naylor, P. Laud, D. Lombard, et al., Essential role of limiting telomeres in the pathogenesis of Werner syndrome, Nat. Genet. 36 (2004) 877–882. [67] C.E. Yu, J. Oshima, Y.H. Fu, E.M. Wijsman, F. Hisama, R. Alisch, et al., Positional cloning of the Werner’s syndrome gene, Science 272 (1996) 258–262. [68] S.E. Artandi, S. Alson, M.K. Tietze, N.E. Sharpless, S. Ye, R.A. Greenberg, et al., Constitutive telomerase expression promotes mammary carcinomas in aging mice, Proc. Natl. Acad. Sci. 99 (2002) 8191–8196. [69] A. Canela, J. Martín-Caballero, J.M. Flores, M.A. Blasco, Constitutive expression of tert in thymocytes leads to increased incidence and dissemination of T-cell lymphoma in Lck-tert mice, Mol. Cell Biol. 24 (2004) 4275–4293. [70] T. Rafnar, P. Sulem, S.N. Stacey, F. Geller, J. Gudmundsson, A. Sigurdsson, et al., Sequence variants at the TERT-CLPTM1L locus associate with many cancer types, Nat. Genet. 41 (2009) 221–227. [71] A. Tomás-Loba, I. Flores, P.J. Fernández-Marcos, M.L. Cayuela, A. Maraver, A. Tejera, et al., Telomerase reverse transcriptase delays aging in cancerresistant mice, Cell 135 (2008) 609–622. [72] B. Bernardes de Jesus, E. Vera, K. Schneeberger, A.M. Tejera, E. Ayuso, F. Bosch, et al., Telomerase gene therapy in adult and old mice delays aging and increases longevity without increasing cancer, EMBO Mol. Med. 4 (2012) 691–704. [73] V. Gorbunova, A. Seluanov, Coevolution of telomerase activity and body mass in mammals: from mice to beavers, Mech. Ageing Dev. 130 (2009) 3–9. [74] N.M.V. Gomes, O.A. Ryder, M.L. Houck, S.J. Charter, W. Walker, N.R. Forsyth, et al., Comparative biology of mammalian telomeres: hypotheses on ancestral states and the roles of telomeres in longevity determination, Aging Cell 10 (2011) 761–768. [75] B.J. Heidinger, J.D. Blount, W. Boner, K. Griffiths, N.B. Metcalfe, P. Monaghan, Telomere length in early life predicts lifespan, Proc. Natl. Acad. Sci. 109 (2012) 1743–1748. [76] R.M. Cawthon, K.R. Smith, E. O’Brien, A. Sivatchenko, R.A. Kerber, Association between telomere length in blood and mortality in people aged 60 years or older, Lancet 361 (2003) 393–395. [77] O.T. Njajou, W.-C. Hsueh, E.H. Blackburn, A.B. Newman, S.-H. Wu, R. Li, et al., Association between telomere length, specific causes of death, and years of healthy life in health, aging, and body composition, a population-based cohort study, J. Gerontol. A: Biol. Sci. Med. Sci. 64A (2009) 860–864. [78] I.M. Wentzensen, L. Mirabello, R.M. Pfeiffer, S.A. Savage, The association of telomere length and cancer: a meta-analysis, Cancer Epidemiol. Biomarkers Prev. 20 (2011) 1238–1250. [79] W. Zhang, Y. Wang, K. Hou, L.-P. Jia, J. Ma, D.-L. Zhao, et al., A correlation study of telomere length in peripheral blood leukocytes and kidney function with age, Mol. Med. Rep. (2015). [80] Z. Li, J. Tang, H. Li, S. Chen, Y. He, Y. Liao, et al., Shorter telomere length in peripheral blood leukocytes is associated with childhood autism, Sci. Rep. (2014) 4. [81] J. Lee, A.J. Sandford, J.E. Connett, J. Yan, T. Mui, Y. Li, et al., The relationship between telomere length and mortality in chronic obstructive pulmonary disease [COPD], PLoS One 7 (2012) e35567. [82] M.J. Machiela, C.A. Hsiung, X.-O. Shu, W.J. Seow, Z. Wang, K. Matsuo, et al., Genetic variants associated with longer telomere length are associated with increased lung cancer risk among never-smoking women in Asia: a report from the female lung cancer consortium in Asia, Int. J. Cancer (2014), n/a –n/a. [83] A.M. Jones, A.D. Beggs, L. Carvajal-Carmona, S. Farrington, A. Tenesa, M. Walker, et al., TERC polymorphisms are associated both with susceptibility to colorectal cancer and with longer telomeres, Gut 61 (2012) 248–254. [84] M.M. Gramatges, M.L. Telli, R. Balise, J.M. Ford, Longer relative telomere length in blood from women with sporadic and familial breast cancer compared with healthy controls, Cancer Epidemiol. Biomarkers Prev. 19 (2010) 605–613. [85] D.C. Cabelof, S. Yanamadala, J.J. Raffoul, Z. Guo, A. Soofi, A.R. Heydari, Caloric restriction promotes genomic stability by induction of base excision repair and reversal of its age-related decline, DNA Repair 2 (2003) 295–307. [86] Z. Guo, A. Heydari, A. Richardson, Nucleotide excision repair of actively transcribed versus nontranscribed DNA in rat hepatocytes: effect of age and dietary restriction, Exp. Cell Res. 245 (1998) 228–238. [87] J.H. Um, S.J. Kim, D.W. Kim, M.Y. Ha, J.H. Jang, D.W. Kim, et al., Tissue-specific changes of DNA repair protein Ku and mtHSP70 in aging rats and their retardation by caloric restriction, Mech. Ageing Dev. 124 (2003) 967–975. [88] H.Y. Cohen, C. Miller, K.J. Bitterman, N.R. Wall, B. Hekking, B. Kessler, et al., Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase, Science 305 (2004) 390–392. [89] A.E. Civitarese, S. Carling, L.K. Heilbronn, M.H. Hulver, B. Ukropcova, W.A. Deutsch, et al., Calorie restriction increases muscle mitochondrial biogenesis in healthy humans, PLoS Med. (2007) 4. [90] J.A. Palacios, D. Herranz, M.L. De Bonis, S. Velasco, M. Serrano, M.A. Blasco, SIRT1 contributes to telomere maintenance and augments global homologous recombination, J. Cell Biol. 191 (2010) 1299–1313. [91] D.E. Harrison, R. Strong, Z.D. Sharp, J.F. Nelson, C.M. Astle, K. Flurkey, et al., Rapamycin fed late in life extends lifespan in genetically heterogenous mice, Nature 460 (2009) 392–395. [92] V. Dao, S. Pandeswara, Y. Liu, V. Hurez, S. Dodds, D. Callaway, et al., Prevention of carcinogen and inflammation-induced dermal cancer by oral rapamycin includes reducing genetic damage, Cancer Prev. Res. 8 (2015) 400–409.

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002

G Model MUT-11549; No. of Pages 7

ARTICLE IN PRESS P. Lenart, L. Krejci / Mutation Research xxx (2016) xxx–xxx

[93] H. Kasai, Analysis of a form of oxidative DNA damage, 8-hydroxy2 -deoxyguanosine, as a marker of cellular oxidative stress during carcinogenesis, Mutat. Res. 387 (1997) 147–163. [94] J.A. Stuart, B.M. Bourque, N.C. de Souza-Pinto, V.A. Bohr, No evidence of mitochondrial respiratory dysfunction in OGG1-null mice deficient in removal of 8-oxodeoxyguanine from mitochondrial DNA, Free Radic. Biol. Med. 38 (2005) 737–745. [95] H. Atamna, I. Cheung, B.N. Ames, A method for detecting abasic sites in living cells: age-dependent changes in base excision repair, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 686–691. [96] M.J. Maclean, R. Aamodt, N. Harris, I. Alseth, E. Seeberg, M. Bjørås, et al., Base excision repair activities required for yeast to attain a full chronological life span, Aging Cell 2 (2003) 93–104. [97] K.W. Caldecott, Single-strand break repair and genetic disease, Nat. Rev. Genet. 9 (2008) 619–631. [98] C.S. Fu, S.B. Harris, P. Wilhelmi, R.L. Walford, Lack of effect of age and dietary restriction on DNA single-stranded breaks in brain, liver, and kidney of [C3H × C57BL/10]F1 Mice, J. Gerontol. 46 (1991) B78–80. [99] A.R. Trzeciak, J.G. Mohanty, K.D. Jacob, J. Barnes, N. Ejiogu, A. Lohani, et al., Oxidative damage to DNA and single strand break repair capacity: relationship to other measures of oxidative stress in a population cohort, Mutat. Res. Mol. Mech. Mutagen. 736 (2012) 93–103. [100] H. Li, J.R. Mitchell, P. Hasty, DNA double-strand breaks: a potential causative factor for mammalian aging, Mech. Ageing Dev. 129 (2008) 416–424. [101] N.P. Singh, C.E. Ogburn, N.S. Wolf, G. van Belle, G.M. Martin, DNA doublestrand breaks in mouse kidney cells with age, Biogerontology 2 (2001) 261–270. [102] O.A. Sedelnikova, I. Horikawa, C. Redon, A. Nakamura, D.B. Zimonjic, N.C. Popescu, et al., Delayed kinetics of DNA double-strand break processing in normal and pathological aging, Aging Cell 7 (2008) 89–100. [103] I.M. Ward, J. Chen, Histone H2AX is phosphorylated in an ATR-dependent manner in response to replicational stress, J. Biol. Chem. 276 (2001) 47759–47762. [104] S. Burma, B.P. Chen, M. Murphy, A. Kurimasa, D.J. Chen, ATM phosphorylates histone H2AX in response to DNA double-strand breaks, J. Biol. Chem. 276 (2001) 42462–42467. [105] J.A. Downs, S. Allard, O. Jobin-Robitaille, A. Javaheri, A. Auger, N. Bouchard, et al., Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites, Mol, Cell 16 (2004) 979–990. [106] E.P. Rogakou, C. Boon, C. Redon, W.M. Bonner, Megabase chromatin domains involved in DNA double-strand breaks in Vivo, J. Cell Biol. 146 (1999) 905–916. [107] Y. Zhang, K. Griffin, N. Mondal, J.D. Parvin, Phosphorylation of histone H2A inhibits transcription on chromatin templates, J. Biol. Chem. 279 (2004) 21866–21872. [108] L.V. Solovjeva, M.P. Svetlova, V.O. Chagin, N.V. Tomilin, Inhibition of transcription at radiation-induced nuclear foci of phosphorylated histone H2AX in mammalian cells, Chromosome Res. 15 (2007) 787–797. [109] R. Di Micco, G. Sulli, M. Dobreva, M. Liontos, O.A. Botrugno, G. Gargiulo, et al., Interplay between oncogene-induced DNA damage response and heterochromatin in senescence and cancer, Nat. Cell Biol. 13 (2011) 292–302. [110] A.A. Goodarzi, P. Jeggo, M. Lobrich, The influence of heterochromatin on DNA double strand break repair: getting the strong, silent type to relax, DNA Repair 9 (2010) 1273–1282. [111] S. Giunta, R. Belotserkovskaya, S.P. Jackson, DNA damage signaling in response to double-strand breaks during mitosis, J. Cell Biol. 190 (2010) 197–207. [112] C.J. Bakkenist, M.B. Kastan, DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation, Nature 421 (2003) 499–506.

7

[113] Z. Bencokova, M.R. Kaufmann, I.M. Pires, P.S. Lecane, A.J. Giaccia, E.M. Hammond, ATM activation and signaling under hypoxic conditions, Mol. Cell Biol. 29 (2009) 526–537. [114] S. Bekker-Jensen, N. Mailand, The ubiquitin- and SUMO-dependent signaling response to DNA double-strand breaks, FEBS Lett. 585 (2011) 2914–2919. [115] L. Moyal, Y. Lerenthal, M. Gana-Weisz, G. Mass, S. So, S.-Y. Wang, et al., Requirement of ATM-dependent monoubiquitylation of histone H2B for timely repair of DNA double-strand breaks, Mol. Cell 41 (2011) 529–542. [116] L. Shi, P. Oberdoerffer, Chromatin dynamics in DNA double-strand break repair, Biochim. Biophys. Acta 1819 (2012) 811–819. [117] B.A. Tamburini, J.K. Tyler, Localized histone acetylation and deacetylation triggered by the homologous recombination pathway of double-strand DNA repair, Mol. Cell Biol. 25 (2005) 4903–4913. [118] M.C. Haigis, D.A. Sinclair, Mammalian Sirtuins: biological insights and disease relevance, Annu. Rev. Pathol. 5 (2010) 253–295. [119] R.I. Tennen, D.J. Bua, W.E. Wright, K.F. Chua, SIRT6 is required for maintenance of telomere position effect in human cells, Nat. Commun. 2 (2011) 433. [120] R. Mostoslavsky, K.F. Chua, D.B. Lombard, W.W. Pang, M.R. Fischer, L. Gellon, et al., Genomic instability and aging-like phenotype in the absence of mammalian SIRT6, Cell 124 (2006) 315–329. [121] C. Chen, A. Jomaa, J. Ortega, E.E. Alani, Pch2 is a hexameric ring ATPase that remodels the chromosome axis protein Hop1, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) E44–E53. [122] H. Qian, X. Xu, L.E. Niklason, PCH-2 regulates Caenorhabditis elegans lifespan, Aging 7 (2015) 1–11. [123] S. Schaetzlein, N. Kodandaramireddy, Z. Ju, A. Lechel, A. Stepzynska, D.R. Lilli, et al., Exo1 deletion impairs DNA damage signal induction and prolongs lifespan of telomere dysfunctional mice, Cell 130 (2007) 863–877. [124] G. Failla, The aging process and cancerogenesis, Ann. N. Y. Acad. Sci. 71 (1958) 1124–1140. [125] L. Szilard, On the nature of the aging process, Proc. Natl. Acad. Sci. U. S. A. 45 (1959) 30–45. [126] M.E.T. Dollé, H. Giese, C.L. Hopkins, H.-J. Martus, J.M. Hausdorff, J. Vijg, Rapid accumulation of genome rearrangements in liver but not in brain of old mice, Nat. Genet. 17 (1997) 431–434. [127] M.E.T. Dollé, W.K. Snyder, J.A. Gossen, P.H.M. Lohman, J. Vijg, Distinct spectra of somatic mutations accumulated with age in mouse heart and small intestine, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 8403–8408. [128] J. Vijg, M.E.T. Dollé, Large genome rearrangements as a primary cause of aging, Mech. Ageing Dev. 123 (2002) 907–915. [129] L. Narayanan, J.A. Fritzell, S.M. Baker, R.M. Liskay, P.M. Glazer, Elevated levels of mutation in multiple tissues of mice deficient in the DNA mismatch repair gene Pms2, Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 3122–3127. [130] R.A. Busuttil, A.M. Garcia, C. Cabrera, A. Rodriquez, Y. Suh, W.H. Kim, et al., Organspecific increase in mutation accumulation and apoptosis rate in CuZn-superoxide dismutase-deficient mice, Cancer Res. 65 (2005) 11271–11275. [131] M.E.T. Dollé, R.A. Busuttil, A.M. Garcia, S. Wijnhoven, E. van Drunen, L.J. Niedernhofer, et al., Increased genomic instability is not a prerequisite for shortened lifespan in DNA repair deficient mice, Mutat. Res. 596 (2006) 22–35. [132] A. Kaya, A.V. Lobanov, V.N. Gladyshev, Evidence that mutation accumulation does not cause aging in Saccharomyces cerevisiae, Aging Cell 14 (2015) 366–371. [133] D.P. Goldman, D. Cutler, J.W. Rowe, P.-C. Michaud, J. Sullivan, D. Peneva, et al., Substantial health and economic returns from delayed aging may warrant a new focus for medical research, Health Aff. [Millwood] 32 (2013) 1698–1705.

Please cite this article in press as: P. Lenart, L. Krejci, Reprint of “DNA, the central molecule of aging”, Mutat. Res.: Fundam. Mol. Mech. Mutagen. (2016), http://dx.doi.org/10.1016/j.mrfmmm.2016.04.002