Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Mar 6.
Published in final edited form as: Nat Cell Biol. 2008 Nov;10(11):1241–1247. doi: 10.1038/ncb1108-1241

DNA damage and ageing: new-age ideas for an age-old problem

George A Garinis 1, Gijsbertus T J van der Horst 2, Jan Vijg 3, Jan H J Hoeijmakers 4
PMCID: PMC4351702  NIHMSID: NIHMS667564  PMID: 18978832

Abstract

Loss of genome maintenance may causally contribute to ageing, as exemplified by the premature appearance of multiple symptoms of ageing in a growing family of human syndromes and in mice with genetic defects in genome maintenance pathways. Recent evidence revealed a similarity between such prematurely ageing mutants and long-lived mice harbouring mutations in growth signalling pathways. At first sight this seems paradoxical as they represent both extremes of ageing yet show a similar ‘survival’ response that is capable of delaying age-related pathology and extending lifespan. Understanding the mechanistic basis of this response and its connection with genome maintenance would open exciting possibilities for counteracting cancer or age-related diseases, and for promoting longevity.


In Greek mythology, Klotho, Lakhesis and Atropos, the three fates, spun, wove and snipped the thread of life, an unalterable process to which both gods and humans had to submit themselves. Human efforts over recent centuries have succeeded in substantially lengthening the thread, allowing ageing to become a common feature of society. However, despite intense research, the molecular basis of the processes that cause loss of bodily functions, and degeneration of cells and tissues is still unresolved. It is widely accepted that ageing is the consequence of stochastic damage accumulation1. Ageing is unique in that it does not seem to be subject to evolutionary selection, as it occurs after the reproductive phase, suggesting that it may occur by default2. Nevertheless, it is apparent from studies in many systems that ageing is subject to regulation by evolutionarily highly conserved molecular pathways35. As such, damage drives functional decline with advancing age; however, the existence of universal mechanisms that are able to promote longevity may set the pace on how rapidly damage builds up and function is lost. We discuss the nature of the processes that determine the length and the quality of the thread of life woven by Lakhesis and ultimately snipped by Atropos.

Damage and ageing: the DNA perspective

Within the complex chemical machinery of each cell, all biomolecules (proteins, lipids and nucleic acids) are subject to indiscriminate damage caused by spontaneous reactions (mostly hydrolysis) and by numerous endogenous and exogenous reactive agents. It is therefore plausible that damage to multiple cellular constituents accounts for ageing1. However, damage to certain macromolecules may play a more prominent part than damage to others. The almost exclusive link between an extending class of syndromes with phenotypes resembling accelerated ageing in many, but not all, organs and tissues (segmental progeria), and inborn defects in DNA metabolism points to genomic damage as a major culprit in the ageing process (Table 1). In principle, all other macromolecules are renewable, whereas nuclear DNA, the blueprint of virtually all cellular RNA and proteins, is irreplaceable; any acquired error is permanent and may have irreversible consequences. In spite of its enormous length and explicit physicochemical vulnerability, cellular function relies on the integrity of the somatic genome, which must be preserved during the entire lifetime of an organism. This is why nature has invested heavily in an intricate genome maintenance apparatus, consisting of several sophisticated DNA damage repair, tolerance and checkpoint systems, as well as effector machinery that enables cell survival or triggers senescence or cell death when DNA is damaged68. This elaborate network also includes intricate machineries to maintain telomeres (the ends of chromosomes), systems to repair mitochondrial DNA and as yet largely unexplored processes that maintain the epigenetic code. These mechanisms ensure that genetic information remains functionally intact for extended periods and is faithfully transmitted. Besides exogenous sources of DNA damage, such as UV and ionizing radiation, and numerous chemicals, there are also inescapable enemies from within. The culprit is the organism’s own metabolism, which generates reactive oxygen species (ROS), including superoxide anions, hydrogen peroxide and hydroxyl radicals and their numerous subsequent reaction products: lipid peroxidation products, oestrogen metabolites, reactive carbonyl species, endogenous alkylating agents, spontaneous hydrolysis and deamination products9. Besides the immense diversity of DNA lesions, the enormous size of the mammalian genome greatly increases its vulnerability to injury, further aggravating the DNA problem. It is estimated that thousands of single-stand breaks and spontaneous base losses occur daily in the nuclear genome of every cell10,11. Together with other types of spontaneous damage, the total may amount to about 100,000 lesions per cell per day11. It is likely that this number increases considerably under certain conditions; for example, a single day in the sun may induce up to 100,000 UV photoproducts in each keratinocyte.

Table 1.

A list of syndromes carrying defects in genome maintenance

Progeria
Syndrome Mutated genes Affected processes Mouse models
Cockayne syndrome (CS) CSA, CSB TC-NER Csa−/−
Csbm/m
TC-NER; GG-NER Csbm/mXpa−/−; Csbm/m/Xpc−/−
Csa−/−Xpa−/−; Csa−/−Xpc−/−
Trichothiodystrophy (TTD) XPB, XPD, TTDA Partial GG/TC-NER Xpdttd
COFS CSB, XPD, XPG GG-NER; TC-NER Xpg−/−
XPE XPF/ERCC1 GG/TC-NER, ICL repair, HR Ercc1−/−
Rothmund-Thomson (RTS) RECQL4 Oxidative DNA damage repair Recql4−/−
Dyskeratosis congenita DKC1, TERC1 Telomere maintenance Dkc1m
mTR−/−
Hutchison-Gilford progeria syndrome (HGPS) LMNA Nuclear lamina function Zmpste24−/−
Atypical Werner syndrome LmnaL530P/L530P
Restrictive dermopathy (RD) LMNA, ZMPSTE24
Mandibuloacral dysplasia (MAD)
Cancer
Syndrome Mutated genes Affected processes Mouse models
Breast cancer 1, early onset BRCA1 DSB repair (HR) Brca1−/−; early lethality
Breast cancer 2, early onset BRAC2 Brca2−/−; early lethality
Li-Fraumeni P53 Checkpoint control p53−/−
Chk2 CHK2 G1 checkpoint control Chk2−/−
von Hippel-Lindau syndrome VHL Cell-cycle regulation Vhl−/−
Hereditary non-polyposis colorectal cancer Msh2;Mlh1 Mismatch repair Msh2−/−
XP XPC GG-NER Xpc−/−
Progeria + cancer
Syndrome Mutated genes Affected processes Mouse models
Fanconi anaemia (FA) FANC, BRCA2 DNA crosslink repair Fancc; Fanca; Fancg; Fancd2;Brca2
Xeroderma pigmentosum (XP) combined with CS (XPCS) XPB, XPF, XPD, XPG NER Xpdxpcs
Xeroderma pigmentosum (XP)+DeSanctis- Cacchione syndrome (DSC) XPA, XPD NER Xpg−/−
Ataxia telangiectasia (AT) ATM DSB repair Atm−/− mTR−/−
Ataxia telangiectasia-like disorder (ATLD) MRE11 DSB repair Mre11−/−
Nijmegen breakage syndrome (NBS) NBS1 DSB and telomere maintenance Nbs1p70
Bloom syndrome (BLS) BLM Mitotic recombination Blm−/−
Werner syndrome (WS) WRN Telomere maintenance, DNA recombination and repair Wrn−/− mTR−/−

Most of the conditions with inborn errors in genome maintenance fall into three classes: 1) those in which many attributes of ageing are accelerated but cancer incidence is reduced, 2) those in which specific cancer types are enhanced and 3) those in which incidence of both cancer and segmental progeria is increased. Mitochondrial heteroplasmic disorders are not shown but should also be considered to belong to the ‘progeria + cancer’ group. Abbreviations: TC-NER: transcription-coupled nucleotide excision repair, COFS: cerebro-oculo-facio-skeletal syndrome, HR: homologous recombination, ICL: interstrand crosslinks, DSB: double-strand break, XFE: Xpf-Errcc1 syndrome.

The consequences of DNA injury are generally unfavourable and determined by various parameters, the first of which is the type of damage. Some lesions are primarily mutagenic, greatly promoting cancer. Others are mainly cytotoxic or cytostatic, triggering cell death or senescence, causing degenerative changes such as those associated with ageing. One of the most common lesions induced by oxidative damage is the highly mutagenic DNA lesion 8-oxo-G, which pairs equally well with the correct cysteine base and the incorrect adenine base, causing high levels of GC to TA transversions. On the other hand, lesions that affect both DNA strands, such as double-strand breaks (induced by radiation or ROS) or interstrand crosslinks (caused by chemical agents such as cisplatin) are difficult to repair. Such lesions usually kill cells or cause senescence rather than mutagenesis (which can lead to tumour formation), possibly accelerating organismal ageing. Indeed, long-term survivors of chemo- or radiotherapy show evidence of premature ageing12,13. Apart from the type of damage, the frequency of lesions and their location in the genome also determine the outcome. Additional parameters are the systems engaged in repairing the damage, notably their fidelity and efficiency, which may vary with cell type and the stage in the cell cycle or differentiation when the damage occurs. As efficient as the genome maintenance machinery may be, it cannot cope with all of the insults inflicted on the genome, leading to a gradual accumulation of DNA damage and mutations14. Certain DNA lesions are poorly, if at all, recognized by the mammalian repair machinery, presumably because they closely resemble the normal DNA conformation. An example is the cyclobutane pyrimidine dimer, the most abundant UV-induced DNA injury, which is often overlooked by global genome repair mechanisms and persists in many parts of the genome15. As this type of damage constitutes a permanent block for the regular high-fidelity replication machinery, a specialized apparatus has developed to allow bypass of persisting damage. This includes a set of at least five translesion polymerases, each specialized in bypassing different types of DNA lesions. Consequently, there may be an elevation in the mutation rate16. Also, dedicated pathways have developed for transcription stalled by DNA damage (see below).

Depending on the type and severity of DNA injuries, failure to repair them, either because they are overlooked or bypassed, may lead to cellular malfunctioning, triggering cancer and senescence, or cell death and eventually loss of organismal homeostasis over time, which contributes to ageing. The wide variety of DNA lesions and their diverse effects have necessitated the development of several layers of protection, including a complementary network of DNA repair pathways, each selective for a specific subset of DNA lesions. The elaborate nature of the genome maintenance apparatus highlights the importance of preserving genome integrity. However, the more complex a system, the more sensitive it is to errors and deficiencies.

Insights from DNA repair-deficient progeroid human syndromes and mouse mutants

Important clues for the clinical effect of DNA damage come from the diverse phenotypes of a rapidly expanding family of rare human disorders associated with genetic defects in DNA repair and damage-response systems. Disorders affecting genome maintenance fall into three classes: 1) conditions in which specific types of cancer are enhanced; 2) conditions in which many (but never all) aspects of ageing are accelerated, but cancer is reduced; 3) conditions in which both cancer and certain aspects of ageing are increased (Table 1). The outcome seems to be governed by the genome maintenance system that is affected. The notion that in none of these syndromes all aspects of cancer and degenerative ageing are equally enhanced is consistent with the idea that each genome stability pathway covers a specific subset of damage and that there are no genes or processes that counteract all forms of DNA damage and their consequences to the same extent.

As is apparent from cancer research over recent decades, analysis of cancer-prone conditions has provided valuable insight into highly relevant pathways for the aetiology of tumorigenesis in general. Similarly, it is expected that the study of progeroid syndromes will highlight molecular mechanisms that normally prevent ageing and age-related disorders17,18. For obvious reasons longevity research generally enjoys wide interest, but strong reservations exist with respect to the relevance of progeroid syndromes and corresponding mouse models of the ageing process19,20. This is largely due to the fact that there are many ways of shortening lifespan, suggesting that premature death alone can indeed be a misleading endpoint. Studies based on lifespan alone, be it extended or shortened, are sensitive to artefacts and it is essential to minimize genetic or environmental sources of variation but also to examine additional ageing parameters, such as evidence of age-related pathologies (as in progeroid models)21. Particularly when this criterion is taken into account, evidence is mounting that DNA damage is a prime, bona fide cause of ageing.

The idea of a double-edged sword of DNA damage — damage-induced mutations causing cancer and damage-triggered cell death/senescence/ malfunction contributing to degenerative forms of ageing22 — is consistent with the phenotypes of DNA repair/genome instability disorders and a growing list of mouse mutants deficient in DNA repair mechanisms. Detailed systematic analysis of these mice, compared with their littermate controls, have revealed the premature appearance of various symptoms of ageing indistinguishable from the same phenotypes normally occurring much later in life. In some cases, a mouse model even paved the way for identifying the parallel human syndrome2326, leaving no doubt that mouse models and human syndromes constitute valid ageing mutants. The overall picture emerging from these mutants is that genetic defects in DNA repair systems that mainly prevent mutagenesis are generally associated with a strong predisposition to specific types of cancer, with only minor symptoms of degenerative ageing phenotypes such as in xeroderma pigmentosum (XP) patients. On the other hand, deficiencies in repair and surveillance pathways that mainly protect from the cytotoxic and cytostatic effects of DNA damage tend to be characterized by a decrease in the incidence of cancer and the premature appearance of some, but not all degenerative ageing phenotypes, such as that of Cockayne syndrome (CS) patients. Impairment of genome stability processes that combat mutagenesis and cell death leads to susceptibility to both cancer and accelerated ageing, such as in patients with both XP and CS (XPCS; Table 1).

An informative example incorporating all aspects discussed above is the nucleotide excision repair (NER) pathway. This is a multi-step ‘detect-excise-and-patch’ repair system for a broad class of helix-distorting lesions such as UV-induced photoproducts and numerous bulky chemical adducts8. Such DNA damage is detected in two ways: 1) the global-genome NER (GG-NER) sub-pathway, which detects lesions with sufficient helix-opening properties anywhere in the genome, and 2) the transcription-coupled NER (TC-NER) sub-pathway, which is selective for lesions that stall the transcription elongation machinery. Both processes detect lesions in a different manner, but the repair process uses the same toolbox, which opens the helix, excises a 22–30-base damage-containing oligonucleotide, fills in the single-strand gap by repair synthesis and ligates the final nick27.

Apart from a common UV hypersensitivity, genetic defects in each of the two sub-pathways have virtually opposite consequences. Impairment of GG-NER in humans causes XP, characterized by an increase of more than 1000-fold in the susceptibility to sun-induced skin cancer28. This is explained by the fact that compromised GG-NER leads to accumulation of DNA lesions over the entire genome and with replication, which increases the risk of mutations. Consistent with this, an increase in the level of mutations is observed, as is apparent from mutagenesis reporter mice29.

On the other hand, genetic defects in TC-NER are associated with the human progeroid disorder CS or the CS-like brittle hair disorder trichothiodystrophy (TTD) that also harbours a partial GG-NER defect. Both conditions and associated mouse models show many symptoms of premature ageing, including progressive neurodevelopmental delay, cachexia, kyphosis, retinal degeneration and deafness28,30,31. Remarkably, TTD mice as well as CS and TTD patients seem to be protected from cancer despite their DNA-repair defect23,28. This is explained by the fact that TC-NER repairs only a small but vital part of the genome, namely the transcribed strand of active genes, when lesions actually block RNA polymerase II32,33. As this system deals with only a tiny fraction of the genome, it is not crucial for preventing mutations and thus cancer. Yet, it is crucial for promoting cell survival after DNA damage, as it enables resumption of the essential process of transcription. Thus, a TC-NER defect increases damage-induced cell death, which prevents damaged cells from surviving and, in effect, this protects from cancer. In a TC-NER mutant, the balance between anti-ageing and anti-cancer genome maintenance responses is shifted to the latter, favouring cell death or senescence, which promotes ageing, while protecting from cancer. Mouse models for these syndromes reveal an absence of significantly elevated spontaneous mutations29, but show markedly accelerated age-related pathology in a number of tissues and organs. This indicates that an increase in point mutations is not a prerequisite for progeria29. Combinations of XP and CS, showing both a predisposition to cancer as well as features of segmental premature ageing, are very rare in patients and are mimicked in the corresponding mouse models34.

XFE is a distinct progeroid syndrome caused by a defect in XPF–ERCC1, an endonuclease required for NER as well as for DNA interstrand crosslink (ICL) repair. ICLs covalently link both strands of DNA, preventing transcription and replication, and hence are extremely cytotoxic. Failing defence against such spontaneous lesions triggers cell death and senescence, culminating in accelerated ageing, as observed in both Ercc1 and Xpf mouse mutants and the human XFE24. Other repair systems, such as base excision repair (eliminating subtle base damages, abasic sites and single-strand breaks) and repair systems for double-strand breaks (homologous recombination and end-joining)8, probably perform both roles; that is, they protect from cancer and ageing to different degrees35,36. Therefore, most defects in distinct DNA repair systems can trigger cancer, ageing or both37, revealing a fine-tuning among genome maintenance mechanisms that mainly protect from cancer, and those that predominantly prevent non-cancer, degenerative ageing phenotypes.

The link between DNA damage and longevity

Genetic crosses in mice have revealed a striking correlation between the severity and type of repair defect and the severity and age of onset of premature ageing features. Crossing progeroid TC-NER-deficient CS and TTD mouse mutants with cancer-prone GG-NER mice (for example, Csbm/m/Xpa−/−, Csa−/−/Xpc−/− and XpdTTD/Xpa−/− mice) substantially aggravates the DNA repair defect (thus increasing the load of endogenous genotoxic stress), further compromising transcription and markedly hastening the onset of progeroid features, including a lifespan reduction from 1.5–2 years to only 3–4 weeks30. As mentioned, Ercc1−/− mice and the recently discovered corresponding human progeroid syndrome XFE24, which are deficient in NER as well as ICL repair, show severe progeroid features that are, in part, different from TC-NER deficiencies (for example dramatic liver, kidney and bone marrow ageing, not observed in TC-NER mutants). Evidently, defects in distinct repair systems for cytotoxic lesions account for a bewildering but specific range of age-related pathologies, which may also explain the distinct segmental nature of progerias.

Recently, the onset of progeroid features in Csbm/m/Xpa−/−, XpdTTD/Xpa−/− and Ercc1−/− repair mutants was shown to be accompanied by marked changes in gene expression and physiological parameters, correlating with a systemic suppression of the growth hormone (GH)/insulin growth factor (IGF)-1 somatotroph axis, suppression of oxidative metabolism, as well as suppression of lactotroph and thyrotroph processes24,30. These changes are paralleled by reduced serum glucose and insulin levels, a consistent upregulation of antioxidant defence and stress responses, along with a marked propensity to store glycogen and fat, indicating an attempt to withhold their energy resources24,30,38. Paradoxically, however, most of these changes, including suppression of the GH/IGF1 hormonal pathway, as well as upregulation of antioxidant and defence responses, are associated with delayed ageing and longevity, as seen in dwarf mutant and calorie-restricted mice, rather than with the extremely short lifespan of NER-deficient progeroid animals39,40.

Indeed, in organisms as evolutionarily diverse as worms and mice, constitutive defects in single genes that perturb endocrine signalling can considerably extend lifespan. Genetic suppression of insulin-signalling in worms prolongs lifespan by several-fold. Similarly, the most prominent pathway affected in long-lived, endocrine-disturbed and dietary-restricted mice is the GH/IGF1 pathway, often paralleled by alterations in thyrotroph and lactotroph functions5,41. For example, Ames and Snell dwarfs42,43, the little mouse (Ghrhrlit/lit)44,45, the homozygous GH receptor/ binding protein (Ghr/bp−/−)46,47, as well as the heterozygous IGF1 receptor (Igf1r+/−)48,49 knockout mice and the Klotho-overexpressing mice50, invariably demonstrate a suppression of the GH/IGF1 somatotroph axis, moderate to pronounced dwarfism and increased lifespan. Similarly, dietary restriction (the only well documented intervention that prolongs lifespan and delays the onset of several ageing-associated diseases in mammals) results in decreased insulin/IGF1 signalling with similar downstream events5156. When dietary restriction and pituitary dwarfism are combined, an additive extension of lifespan is observed57. Thus, suppression of the GH/IGF1 axis is associated with delayed age-related morbidity and longevity, profound metabolic changes (including low serum glucose and insulin), enhanced antioxidant defences and stress resistance, and reduced frequency of somatic mutations48,58,59. Conversely, overexpression of GH causes pathology and markedly shortens lifespan60. The genome-wide expression parallels between long-lived mutant dwarf or CR mice and the NER progeroid mutants40 explain the arrested growth and development of NER progeria, which are not caused by a defect in the hypothalamus or the pituitary of NER progeroid mutants24.

Studies on dwarf mutants suggest that lifespan can, in principle, be genetically modulated by single gene mutations, although ageing is not abolished altogether by these mutations. CR interventions later in life demonstrate that longevity assurance mechanisms can be adaptive and to some extent reversible61. How does this relate to the irreversible, random damage theory of ageing62? Given the abundance of evidence for both models, is there a link between conserved pathways that promote longevity by suppressing the random accumulation of damage? According to the ‘disposable soma’ theory, ageing arises from the accumulation of macromolecular damage because of inherent limitations in somatic maintenance and repair1. As energy reserves that are engaged in one process are unavailable to contribute to another, it is conceivable that an organism can only maximize its Darwinian fitness by optimally allocating its metabolic needs between the maintenance and repair of its soma and physiological processes, such as reproduction and development. Under affluent conditions it would give preference to rapid reproduction, whereas in times of scarcity, reproduction would be postponed and a larger share of resources devoted to maintenance. The NER progeroid mice could resemble the stress condition normally associated with food restriction and elicit a very similar adaptive response. It is appealing to speculate that the progressive changes associated with the insulin/IGF1 pathway and metabolism are adaptive responses aimed at minimizing further damage, by shifting the energy equilibrium from growth and proliferation to preservation of somatic maintenance (Fig. 1). This adaptive ‘survival’ response is likely to be driven by intrinsic genome instability, but it can also be triggered by other emergency situations such as calorie restriction. In parallel, accumulation of DNA damage may trigger similar responses with normal ageing. Indeed, chronic exposure of wild-type mice to non-toxic doses of pro-oxidant or crosslinking agents elicits the same GH/IGF1 suppression63. Thus, the gradual accumulation of damage could also explain the known GH/IGF1 somatotroph attenuation in naturally aged mammals. In support of this, Csbm/m/Xpa−/− and Ercc1−/− mice demonstrate significant, genome-wide expression parallels with naturally aged mice, indicating that stochastic accumulation of damage to macromolecules (including DNA) may cause the physiological decline in the GH/IGF1 axis and organismal deterioration with advancing age24,30.

Figure 1.

Figure 1

Schematic representation of ‘stochastic damage’ and the connection with ageing and longevity assurance mechanisms. DNA is continually damaged by chemical alterations (such as spontaneous hydrolysis, deamination), by environmental agents as well as endogenous products (that is, ROS). Cells respond through a battery of DNA repair and genome surveillance systems that counteract DNA damage, thereby ensuring that their vital genetic information is preserved and faithfully transmitted to progeny. Nevertheless, a fraction of the damage escapes repair and accumulates, resulting in mutations, senescence or cell death and cellular dysfunction. Too much persisting DNA damage interferes with normal DNA metabolism, such as transcription, and triggers suppression of the growth hormone/IGF1 somatotropic axis, which is known to decline with age. Dampening of the insulin/IGF1 pathway and oxidative metabolism is thought to reduce the induction and effects of DNA damage by shifting the energy equilibrium from growth and proliferation to pathways that preserve somatic maintenance and thus attempt to extend lifespan (survival response). NER progeroid mice accumulate DNA damage much more rapidly than naturally ageing mice as a consequence of their repair defect, and onset of the life-extending ‘survival’ response is accelerated. Thus, studies in mice with inherited defects in genome maintenance seem to reconcile two apparently contrasting theories on ageing: the genetic basis of ageing and the stochastic damage accumulation. As random damage drives the age-related functional decline, longevity assurance mechanisms determine the rate of damage accumulation and the functional decline with age.

As yet we do not know the sequence of events leading from DNA damage to the activation of longevity assurance pathways. It is likely that there are other protective mechanisms that will postpone cancer as well as ageing. These include reduced or better controlled production of metabolic byproducts through regulation of metabolism and oxidative phosphorylation, as well as improved anti-oxidant/detoxification defences, which reduce the DNA damage load and its noxious sequelae. Many of these systems seem to be under the control of the highly conserved GH/IGF1 somatotroph axis and are triggered by the survival response. In addition, improved overall repair pathways of both categories, that is, anti-cancer and anti-ageing, are predicted to extend lifespan next to other mechanisms, including epimutations and protein metabolism, which are relevant for both ageing and cancer. Presumably, gradual evolutionary improvements in these systems have enabled longer-lived species, such as humans, to achieve their long lifespan by combating cancer and ageing simultaneously. This model also holds the promise that we may find compounds that either trigger the protective survival response or diminish the oxidative (DNA) damage load by effectively scavenging deleterious ROS. Promising indications in these directions have already emerged, such as the design and synthesis of small dietary restriction mimetics that could potentially promote healthy and long lifespan in higher organisms64.

Acknowledgments

This work was supported by the Netherlands Organization for Scientific Research (NWO) through the foundation of the Research Institute Diseases of the Elderly, as well as grants from SenterNovem IOP-Genomics (IGE03009), NIH (1PO1 AG17242-02), NIEHS (1UO1 ES011044), EC (QRTL-1999-02002), and the Dutch Cancer Society (EUR 99-2004).

Contributor Information

George A. Garinis, Department of Genetics, Centre for Biomedical Genetics, Erasmus University Medical Centre and is now at the Institute of Molecular Biology and Biotechnology, FORTH, Vassilika Vouton, P.O. Box 1385, GR 711 10 Heraklion, Crete, Greece

Gijsbertus T. J. van der Horst, Department of Genetics, Centre for Biomedical Genetics, Erasmus University Medical, Centre, PO Box 2040, 3000 CA Rotterdam, the Netherlands

Jan Vijg, Buck Institute for Aging Research, 8001 Redwood Blvd, Novato, CA 94945, USA and is now in the Department of Genetics, Albert Einstein College of Medicine, 1301 Morris Park Avenue, Bronx, NY 10461-1926, USA.

Jan H. J. Hoeijmakers, Email: j.hoeijmakers@erasmusmc.nl, Department of Genetics, Centre for Biomedical Genetics, Erasmus University Medical, Centre, PO Box 2040, 3000 CA Rotterdam, the Netherlands

References

  • 1.Kirkwood TB. Understanding the odd science of aging. Cell. 2005;120:437–447. doi: 10.1016/j.cell.2005.01.027. [DOI] [PubMed] [Google Scholar]
  • 2.Kirkwood TB, Cremer T. Cytogerontology since 1881: a reappraisal of August Weismann and a review of modern progress. Hum Genet. 1982;60:101–121. doi: 10.1007/BF00569695. [DOI] [PubMed] [Google Scholar]
  • 3.Guarente L, Kenyon C. Genetic pathways that regulate ageing in model organisms. Nature. 2000;408:255–262. doi: 10.1038/35041700. [DOI] [PubMed] [Google Scholar]
  • 4.Partridge L, Gems D. Mechanisms of ageing: public or private? Nature Rev Genet. 2002;3:165–175. doi: 10.1038/nrg753. [DOI] [PubMed] [Google Scholar]
  • 5.Kenyon C. The plasticity of aging: insights from long-lived mutants. Cell. 2005;120:449–460. doi: 10.1016/j.cell.2005.02.002. [DOI] [PubMed] [Google Scholar]
  • 6.d’Adda di Fagagna F, Teo SH, Jackson SP. Functional links between telomeres and proteins of the DNA-damage response. Genes Dev. 2004;18:1781–1799. doi: 10.1101/gad.1214504. [DOI] [PubMed] [Google Scholar]
  • 7.Harper JW, Elledge SJ. The DNA damage response: ten years after. Mol Cell. 2007;28:739–745. doi: 10.1016/j.molcel.2007.11.015. [DOI] [PubMed] [Google Scholar]
  • 8.Hoeijmakers JH. Genome maintenance mechanisms for preventing cancer. Nature. 2001;411:366–374. doi: 10.1038/35077232. [DOI] [PubMed] [Google Scholar]
  • 9.De Bont R, van Larebeke N. Endogenous DNA damage in humans: a review of quantitative data. Mutagenesis. 2004;19:169–185. doi: 10.1093/mutage/geh025. [DOI] [PubMed] [Google Scholar]
  • 10.Sander M, et al. Proceedings of a workshop on DNA adducts: biological significance and applications to risk assessment Washington, DC, April 13–14, 2004. Toxicol Appl Pharmacol. 2005;208:1–20. doi: 10.1016/j.taap.2004.12.012. [DOI] [PubMed] [Google Scholar]
  • 11.Lindahl T. Instability and decay of the primary structure of DNA. Nature. 1993;362:709–715. doi: 10.1038/362709a0. [DOI] [PubMed] [Google Scholar]
  • 12.Grillari J, Katinger H, Voglauer R. Contributions of DNA interstrand cross-links to aging of cells and organisms. Nucleic Acids Res. 2007;35:7566–7576. doi: 10.1093/nar/gkm1065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beeharry N, Broccoli D. Telomere dynamics in response to chemotherapy. Curr Mol Med. 2005;5:187–196. doi: 10.2174/1566524053586554. [DOI] [PubMed] [Google Scholar]
  • 14.Vijg J. Somatic mutations and aging: a re-evaluation. Mutat Res. 2000;447:117–135. doi: 10.1016/s0027-5107(99)00202-x. [DOI] [PubMed] [Google Scholar]
  • 15.Friedberg EC, WG, Siede W, Wood RD, Schultz RA, Ellenberger T. DNA Repair and Mutagenesis 2nd edn. ASM; Washington DC: 2006. [Google Scholar]
  • 16.Plosky BS, Woodgate R. Switching from high-fidelity replicases to low-fidelity lesion-bypass polymerases. Curr Opin Genet Dev. 2004;14:113–119. doi: 10.1016/j.gde.2004.02.002. [DOI] [PubMed] [Google Scholar]
  • 17.Martin GM. Genetic modulation of senescent phenotypes in Homo sapiens. Cell. 2005;120:523–532. doi: 10.1016/j.cell.2005.01.031. [DOI] [PubMed] [Google Scholar]
  • 18.Vilhem Bohr DW, de Souza Pinto NC, van der Pluijm I, Hoeijmakers JH. DNA Repair and Aging. Cold Spring Harbor Laboratory Press; New York: 2008. [Google Scholar]
  • 19.Hasty P, Vijg J. Rebuttal to Miller: ‘Accelerated aging’: a primrose path to insight?’. Aging Cell. 2004;3:67–69. doi: 10.1111/j.1474-9728.2004.00087.x. [DOI] [PubMed] [Google Scholar]
  • 20.Miller RA. ‘Accelerated aging’: a primrose path to insight? Aging Cell. 2004;3:47–51. doi: 10.1111/j.1474-9728.2004.00081.x. [DOI] [PubMed] [Google Scholar]
  • 21.Partridge L, Gems D. Benchmarks for ageing studies. Nature. 2007;450:165–167. doi: 10.1038/450165a. [DOI] [PubMed] [Google Scholar]
  • 22.Campisi J. Senescent cells, tumor suppression, and organismal aging: good citizens, bad neighbors. Cell. 2005;120:513–522. doi: 10.1016/j.cell.2005.02.003. [DOI] [PubMed] [Google Scholar]
  • 23.Wijnhoven SW, et al. Accelerated aging pathology in ad libitum fed Xpd(TTD) mice is accompanied by features suggestive of caloric restriction. DNA Repair. 2005;4:1314–1324. doi: 10.1016/j.dnarep.2005.07.002. [DOI] [PubMed] [Google Scholar]
  • 24.Niedernhofer LJ, et al. A new progeroid syndrome reveals that genotoxic stress suppresses the somatotroph axis. Nature. 2006;444:1038–1043. doi: 10.1038/nature05456. [DOI] [PubMed] [Google Scholar]
  • 25.de Boer J, et al. Premature aging in mice deficient in DNA repair and transcription. Science. 2002;296:1276–1279. doi: 10.1126/science.1070174. [DOI] [PubMed] [Google Scholar]
  • 26.Hasty P, Campisi J, Hoeijmakers J, van Steeg H, Vijg J. Aging and genome maintenance: lessons from the mouse? Science. 2003;299:1355–1359. doi: 10.1126/science.1079161. [DOI] [PubMed] [Google Scholar]
  • 27.Scharer OD. A molecular basis for damage recognition in eukaryotic nucleotide excision repair. Chembiochem. 2008;9:21–23. doi: 10.1002/cbic.200700619. [DOI] [PubMed] [Google Scholar]
  • 28.Bootsma DK, KH, Cleaver JE, Hoeijmakers JHJ. The Metabolic and Molecular Basis of inherited Disease. McGraw-Hill; New York: 2001. pp. 677–703. [Google Scholar]
  • 29.Dolle ME, et al. Increased genomic instability is not a prerequisite for shortened lifespan in DNA repair deficient mice. Mutat Res. 2006;596:22–35. doi: 10.1016/j.mrfmmm.2005.11.008. [DOI] [PubMed] [Google Scholar]
  • 30.van der Pluijm I, et al. Impaired genome maintenance suppresses the growth hormone–insulin-like growth factor 1 axis in mice with Cockayne syndrome. PLoS Biol. 2006;5:e2. doi: 10.1371/journal.pbio.0050002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.de Boer J, et al. Mouse model for the DNA repair/basal transcription disorder trichothiodystrophy reveals cancer predisposition. Cancer Res. 1999;59:3489–3494. [PubMed] [Google Scholar]
  • 32.Hanawalt PC. Subpathways of nucleotide excision repair and their regulation. Oncogene. 2002;21:8949–8956. doi: 10.1038/sj.onc.1206096. [DOI] [PubMed] [Google Scholar]
  • 33.Ljungman M, Lane DP. Transcription — guarding the genome by sensing DNA damage. Nature Rev Cancer. 2004;4:727–737. doi: 10.1038/nrc1435. [DOI] [PubMed] [Google Scholar]
  • 34.Andressoo JO, Hoeijmakers JH, Mitchell JR. Nucleotide excision repair disorders and the balance between cancer and aging. Cell Cycle. 2006;5:2886–2888. doi: 10.4161/cc.5.24.3565. [DOI] [PubMed] [Google Scholar]
  • 35.Li H, Vogel H, Holcomb VB, Gu Y, Hasty P. Deletion of Ku70, Ku80, or both causes early aging without substantially increased cancer. Mol Cell Biol. 2007;27:8205–8214. doi: 10.1128/MCB.00785-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Vogel H, Lim DS, Karsenty G, Finegold M, Hasty P. Deletion of Ku86 causes early onset of senescence in mice. Proc Natl Acad Sci USA. 1999;96:10770–10775. doi: 10.1073/pnas.96.19.10770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Andressoo JO, et al. An Xpd mouse model for the combined xeroderma pigmentosum/ Cockayne syndrome exhibiting both cancer predisposition and segmental progeria. Cancer Cell. 2006;10:121–132. doi: 10.1016/j.ccr.2006.05.027. [DOI] [PubMed] [Google Scholar]
  • 38.van de Ven M, et al. Adaptive stress response in segmental progeria resembles long-lived dwarfism and calorie restriction in mice. PLoS Genet. 2006;2:e192. doi: 10.1371/journal.pgen.0020192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Schumacher B, Garinis GA, Hoeijmakers JH. Age to survive: DNA damage and aging. Trends Genet. 2008;24:77–85. doi: 10.1016/j.tig.2007.11.004. [DOI] [PubMed] [Google Scholar]
  • 40.Schumacher B, et al. Delayed and accelerated aging share common longevity assurance mechanisms. PLoS Genet. 2008;4:e1000161. doi: 10.1371/journal.pgen.1000161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bartke A. Minireview: role of the growth hormone/insulin-like growth factor system in mammalian aging. Endocrinology. 2005;146:3718–3723. doi: 10.1210/en.2005-0411. [DOI] [PubMed] [Google Scholar]
  • 42.Li S, et al. Dwarf locus mutants lacking three pituitary cell types result from mutations in the POU-domain gene pit-1. Nature. 1990;347:528–533. doi: 10.1038/347528a0. [DOI] [PubMed] [Google Scholar]
  • 43.Andersen B, et al. The Ames dwarf gene is required for Pit-1 gene activation. Dev Biol. 1995;172:495–503. doi: 10.1006/dbio.1995.8040. [DOI] [PubMed] [Google Scholar]
  • 44.Godfrey P, et al. GHRH receptor of little mice contains a missense mutation in the extracellular domain that disrupts receptor function. Nature Genet. 1993;4:227–232. doi: 10.1038/ng0793-227. [DOI] [PubMed] [Google Scholar]
  • 45.Flurkey K, Papaconstantinou J, Miller RA, Harrison DE. Lifespan extension and delayed immune and collagen aging in mutant mice with defects in growth hormone production. Proc Natl Acad Sci USA. 2001;98:6736–6741. doi: 10.1073/pnas.111158898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zhou Y, et al. A mammalian model for Laron syndrome produced by targeted disruption of the mouse growth hormone receptor/binding protein gene (the Laron mouse) Proc Natl Acad Sci USA. 1997;94:13215–13220. doi: 10.1073/pnas.94.24.13215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Coschigano KT, Clemmons D, Bellush LL, Kopchick JJ. Assessment of growth parameters and life span of GHR/BP gene-disrupted mice. Endocrinology. 2000;141:2608–2613. doi: 10.1210/endo.141.7.7586. [DOI] [PubMed] [Google Scholar]
  • 48.Holzenberger M, et al. IGF-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature. 2003;421:182–187. doi: 10.1038/nature01298. [DOI] [PubMed] [Google Scholar]
  • 49.Liu JP, Baker J, Perkins AS, Robertson EJ, Efstratiadis A. Mice carrying null mutations of the genes encoding insulin-like growth factor I (Igf-1) and type 1 IGF receptor (Igf1r) Cell. 1993;75:59–72. [PubMed] [Google Scholar]
  • 50.Kurosu H, et al. Suppression of aging in mice by the hormone Klotho. Science. 2005;309:1829–1833. doi: 10.1126/science.1112766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.McCay CM, Crowell MF, Maynard LA. The effect of retarded growth upon the length of life span and upon the ultimate body size. 1935. Nutrition. 1989;5:155–171. [PubMed] [Google Scholar]
  • 52.Maeda H, et al. Nutritional influences on aging of Fischer 344 rats: II. Pathology. J Gerontol. 1985;40:671–688. doi: 10.1093/geronj/40.6.671. [DOI] [PubMed] [Google Scholar]
  • 53.Masoro EJ. Dietary restriction and aging. J Am Geriatr Soc. 1993;41:994–999. doi: 10.1111/j.1532-5415.1993.tb06767.x. [DOI] [PubMed] [Google Scholar]
  • 54.Masoro EJ. Food restriction in rodents: an evaluation of its role in the study of aging. J Gerontol. 1988;43:B59–B64. doi: 10.1093/geronj/43.3.b59. [DOI] [PubMed] [Google Scholar]
  • 55.Miller RA, et al. Gene expression patterns in calorically restricted mice: partial overlap with long-lived mutant mice. Mol Endocrinol. 2002;16:2657–2666. doi: 10.1210/me.2002-0142. [DOI] [PubMed] [Google Scholar]
  • 56.Bauer M, et al. Starvation response in mouse liver shows strong correlation with lifespan-prolonging processes. Physiol Genomics. 2004;17:230–244. doi: 10.1152/physiolgenomics.00203.2003. [DOI] [PubMed] [Google Scholar]
  • 57.Tsuchiya T, et al. Additive regulation of hepatic gene expression by dwarfism and caloric restriction. Physiol Genomics. 2004;17:307–315. doi: 10.1152/physiolgenomics.00039.2004. [DOI] [PubMed] [Google Scholar]
  • 58.Garcia AM, et al. Effect of Ames dwarfism and caloric restriction on spontaneous DNA mutation frequency in different mouse tissues. Mech Ageing Dev. 2008;129:528–533. doi: 10.1016/j.mad.2008.04.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Liang H, et al. Genetic mouse models of extended lifespan. Exp Gerontol. 2003;38:1353–1364. doi: 10.1016/j.exger.2003.10.019. [DOI] [PubMed] [Google Scholar]
  • 60.Bartke A, et al. Effects of growth hormone overexpression and growth hormone resistance on neuroendocrine and reproductive functions in transgenic and knock-out mice. Proc Soc Exp Biol Med. 1999;222:113–123. doi: 10.1046/j.1525-1373.1999.d01-121.x. [DOI] [PubMed] [Google Scholar]
  • 61.Cao SX, Dhahbi JM, Mote PL, Spindler SR. Genomic profiling of short- and long-term caloric restriction effects in the liver of aging mice. Proc Natl Acad Sci USA. 2001;98:10630–10635. doi: 10.1073/pnas.191313598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Harman D. The aging process. Proc Natl Acad Sci USA. 1981;78:7124–7128. doi: 10.1073/pnas.78.11.7124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.van der Pluijm IGG, et al. Impaired genome maintenance suppresses the growth hormone–insulin-like growth factor 1 axis in mice with cockayne syndrome. PLoS Biol. 2006;5:23–38. doi: 10.1371/journal.pbio.0050002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Yang H, Baur JA, Chen A, Miller C, Sinclair DA. Design and synthesis of compounds that extend yeast replicative lifespan. Aging Cell. 2007;6:35–43. doi: 10.1111/j.1474-9726.2006.00259.x. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES