Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Oct 22.
Published in final edited form as: Mol Cell. 2010 Oct 22;40(2):333–344. doi: 10.1016/j.molcel.2010.10.002

The Aging Stress Response

Marcia C Haigis 1, Bruce A Yankner 1
PMCID: PMC2987618  NIHMSID: NIHMS247203  PMID: 20965426

Abstract

Aging is the outcome of a balance between damage and repair. The rate of aging and the appearance of age-related pathology are modulated by stress response and repair pathways that gradually decline, including the proteostasis and DNA damage repair networks and mitochondrial respiratory metabolism. Highly conserved insulin/IGF-1, TOR, and sirtuin signaling pathways in turn, control these critical cellular responses. The coordinated action of these signaling pathways maintains cellular and organismal homeostasis in the face of external perturbations, such as changes in nutrient availability, temperature and oxygen level, as well as internal perturbations, such as protein misfolding and DNA damage. Studies in model organisms suggest that changes in signaling can augment these critical stress response systems, increasing lifespan and reducing age-related pathology. The systems biology of stress response signaling thus provides a new approach to the understanding and potential treatment of age-related diseases.

Keywords: aging, DNA damage, mitochondria, stress response, proteostasis, autophagy, insulin, TOR, sirtuin

Introduction

A phylogenetically conserved feature of aging is the induction of stress response pathways. Microarray studies of gene expression show that age-dependent induction of stress response genes occurs in all systems studied, including whole organism analysis of Caenorhabditis elegans (C. elegans) and Drosophila, and analysis of the brain in mouse, rat, chimpanzee and human (Bishop et al., 2010; Yankner et al., 2008). Cellular stress response pathways are controlled at the molecular level by a number of highly conserved signaling molecules and transcriptional regulators, including proteins involved in insulin/insulin-like growth factor (IGF) signaling, sirtuins, target of rapamycin (TOR) and AMP-activated protein kinase (AMPK) pathways (Kenyon, 2010). The molecules involved possess features enabling them to both sense changes in inputs, such as energy status, DNA damage, protein damage, and hypoxia, and transmit information to molecules that allow for adaptive cellular responses. These adaptive responses often involve coordinated regulation of protein synthesis and turnover, autophagy, and mitochondrial function. This highly regulated circuitry thus maintains protein and DNA integrity in the face of stress and declining function during aging.

This review explores the connection between stress response pathways and aging. We begin by discussing the evidence that aging is a regulated process that is controlled by a few highly conserved signaling mechanisms, including the insulin/IGF-1, TOR, and AMPK pathways, as well as sirtuins. Evidence is discussed for coordinated regulation by these signaling pathways of stress responses that play a role in aging, including nutrient sensing (Sengupta, 2010), mitochondrial function, redox metabolism (Majmundar et al., 2010; Wellen et al., 2010), the DNA damage response (Ciccia, 2010), proteostasis (Buchberger, 2010; Richter, 2010) and autophagy (Kroemer, 2010). The potential for harnessing the phenomenon of hormesis, in which low level stressors enhance organismal resistance and reduce physiological decline, is described as an emerging concept. Recent findings suggest that stress response signaling can be manipulated to extend lifespan and retard the onset of age-related pathology, suggesting a new approach to the degenerative disorders of aging humans.

Nutrient Signaling Pathways and Aging

Lifespan is regulated by highly conserved nutrient sensing pathways providing evidence for a pivotal role of nutrient signaling in the control of aging and aging-related diseases. High caloric intake shortens lifespan and accelerates the onset of aging-associated disorders, including diabetes, metabolic syndrome, cancer and neurodegenerative disorders. By contrast, a dietary regimen of moderate calorie restriction with adequate nutrient intake delays aging in a wide variety of organisms from yeast to primates, and may delay or attenuate age-related diseases such as diabetes, cancer and Alzheimer’s disease (AD) (Comfort, 1963; Haigis and Sinclair, 2010; McCay et al., 1989; Weindruch and Walford, 1988). Caloric restriction also activates stress pathways that increase organismal resistance to subsequent stress or nutritional limitation, an effect known as hormesis. Energy sensing pathways are linked to the aging process and are regulated by insulin/IGF-1, sirtuins, TOR, and AMPK signaling (Kenyon, 2010). Moreover, recent studies demonstrate that these pathways coordinately regulate each other, as well as a variety of stress response pathways that impact organismal survival and lifespan. Understanding how organisms sense nutrient intake and stress and coordinate these signals is likely to increase our understanding of mechanisms that underlie aging and age-related diseases.

Insulin/IGF-1 signaling pathway

Mutations that reduce insulin/IGF-1 signaling extend lifespan in a variety of model organisms. In C. elegans, loss of function mutations in daf-2, a homolog of mammalian insulin/ IGF receptors, extends lifespan by more than 2-fold (Kenyon et al., 1993; Kimura et al., 1997). The regulation of aging by insulin-like factors involves downstream signaling through phosphatidylinositol 3-OH kinase (PI(3)K), AKT and FOXOs; in worms, the AGE-1 mutation in an insulin/IGF-regulated PI(3)K ortholog increases lifespan (Morris et al., 1996). Furthermore, lifespan extension requires signaling through the transcription factors DAF-16, HSF-1 (Heat Shock Factor-1), and SKN-1 that induce the expression of a broad network of genes involved in anti-oxidant defense, mitochondrial function, proteostasis and autophagy (Kenyon, 2010). Thus, signaling through insulin/IGF-1 modulates cellular and organismal stress responses by controlling key transcriptional programs.

Because insulin/ signaling functions as a nutrient sensor and controls transcription of stress response genes, this pathway provides a molecular connection between dietary intake and cellular stress response pathways. Indeed, experimental data support that idea that insulin/IGF-1 pathways mediate at least part of the beneficial effect of calorie restriction. For example, caloric restriction decreases the levels of insulin/IGF-1 in mammals and may be perceived by an organism as a mild type of stress, providing hormetic benefits (Comfort, 1963; Haigis and Sinclair, 2010; McCay et al., 1989; Weindruch and Walford, 1988). Interestingly, DAF-16 and SKN1, which are regulated by insulin/IGF-1 signaling, are required for lifespan extension in some models of calorie restriction in worms (Bishop and Guarente, 2007; Greer and Brunet, 2009). In flies, the lifespan extension by dietary conditions seems to be mediated, in part, by FOXO activity (Giannakou et al., 2008). The insulin/IGF-1 signaling pathway also appears to play a central role in the beneficial effects of caloric restriction in mammals. Growth hormone receptor knockout mice are long-lived, and their lifespan cannot be further extended by dietary restriction (Bonkowski et al., 2009), suggesting that reduced IGF-1 signaling and caloric restriction extend lifespan through similar mechanisms.

Recent genetic studies have found inherited SNPs in genes of the insulin/IGF-1 signaling pathway that correlate with longevity. Polymorphisms in the IGF-1 receptor gene have been identified in Ashkenazi Jewish centenarians (Suh et al., 2008). SNPS have also been identified in the insulin signaling genes AKT1, FOXO1 and FOXO3a in multiple centenarian cohorts (Wilcox et al., 2008;Flachsbart et al., 2009; Pawlikowska et al., 2009). Larger population based studies will be required to determine whether these genetic associations represent true human longevity traits.

Sirtuins

Sirtuins are a highly conserved family of proteins that connect metabolic status to the regulation of aging and age-related phenotypes (Haigis and Sinclair, 2010). Sirtuins possess NAD-dependent protein deacetylase and/or ADP-ribosyltransferase activity. The requirement for NAD is one mechanism by which sirtuins sense and respond to metabolic status (Guarente, 2006; Schwer and Verdin, 2008). A number of studies in model organisms, including yeast, worms and flies, suggest that the sirtuin Sir2 can extend lifespan (Kaeberlein et al., 1999; Lin et al., 2000) (Rogina and Helfand, 2004; Tissenbaum and Guarente, 2001). There are seven mammalian sirtuins (SIRT1-7) that play various roles in the regulation of stress resistance, metabolism and cell survival. However, their roles in the regulation of mammalian lifespan are still unresolved. Nonetheless, many reports suggest that sirtuins regulate stress-response pathways that contribute to aging and age-related diseases.

The best studied mammalian sirtuin is SIRT1, the mammalian ortholog of yeast Sir2 (Bordone and Guarente, 2005; Frye, 2000). SIRT1 activity is regulated by cellular nutrient status and triggers stress response pathways and changes in energy metabolism. Moreover, SIRT1 expression and activity decrease with age in a number of tissues, and can also be reduced by a high fat diet (Ramsey et al., 2008; Sasaki et al., 2006). Conversely, SIRT1 activity is increased during times of nutrient deprivation, such as fasting and calorie restriction. When activated, SIRT1 deacetylates many different substrate proteins that are involved in aging, stress responses and metabolic regulation, including PGC-1α, Ku70, NF-κB, AceCS1, MEF2 and p53 (Haigis and Sinclair, 2010) Moreover, SIRT1 deacetylates and activates FOXO transcription factors, providing a level of transcriptional control of stress response genes.

SIRT1-deficient mice develop insulin resistance and metabolic deficits that may relate, in part, to impaired energy metabolism (Haigis and Sinclair, 2010). Mitochondria isolated from SIRT1-deficient mice show reduced respiratory function and increased generation of reactive oxygen species (ROS) (Boily et al., 2008). By contrast, mice overexpressing SIRT1 show improved metabolic parameters that resemble the metabolic changes associated with caloric restriction (Banks et al., 2008; Bordone et al., 2007). Moreover, activation of SIRT1 by resveratrol may contribute to lifespan extension in mice fed a high fat diet (Baur et al., 2006; Lagouge et al., 2006). However, overexpression of SIRT1 in transgenic mice fed a normal diet does not extend lifespan (Herranz et al., 2010), raising the possibility that SIRT1 function is most relevant to lifespan regulation under stress-related conditions in mammals. In addition to SIRT1, studies of the other sirtuins implicate this family of proteins in the regulation of multiples aspects of stress resistance and metabolism, including DNA repair, mitochondrial function, protein quality control, and cell survival (Haigis and Sinclair, 2010). Taken together, these studies suggest that sirtuins activate protective stress responses in a variety of model systems, but the role of sirtuins in the regulation of mammalian lifespan remains to be determined.

AMPK

AMP-activated protein kinase (AMPK) is activated under conditions of elevated intracellular AMP or reduced ATP, enabling this kinase to serve as a rheostat for cellular energy status. Stressors such as glucose deprivation, ischemia, hypoxia and exercise that deplete cellular ATP lead to the activation of AMPK (Kahn et al., 2005). AMPK activation results in transcriptional and post-translational signaling responses that increase catabolic metabolic pathways in response to the stress of a low energy state (Nilsson et al., 2006; Osler and Zierath, 2008). For example, AMPK activity promotes fatty acid oxidation through phosphorylation and inhibition of ACC, an enzyme that synthesizes malonyl CoA from acetyl CoA. The regulation of ACC activity is a pivotal node in the switch between anabolic and catabolic processes. The malonyl CoA generated by ACC provides a precursor for fatty acid synthesis, while also inhibiting mitochondrial fatty acid oxidation via allosteric inhibition of carnitine palmitoyltransferase-1 (CPT1), the rate-limiting enzyme in mitochondrial fatty acid uptake. (Abu-Elheiga et al., 2001; Saggerson, 2008). Thus, during periods of nutrient stress resulting in low ATP, the switch from fatty acid synthesis to oxidation is mediated in a large part by the increased activity of AMPK.

Declining AMPK activity during aging may contribute to insulin resistance and metabolic syndrome. AMPK activity declines in aging skeletal muscle, and is associated with insulin resistance that can be reversed by treatment with AICAR, an AMP analog that activates AMPK (Qiang et al., 2007). AMPK may regulate insulin sensitivity by stimulating GLUT4 translocation, increasing glucose uptake and metabolism. These observations suggest that AMPK is a potential therapeutic target for age-related metabolic disorders (McCarty, 2004).

mTOR

The target of rapamycin (TOR) pathway is a conserved nutrient sensor that is linked to lifespan regulation. TOR integrates environmental cues, such as growth factors and nutrients to control eukaryotic growth, metabolism and cell division (Cunningham et al., 2007; Wullschleger et al., 2006). TOR activity is suppressed by conditions of nutrient limitation, consistent with the notion that decreasing TOR signaling mimics aspects of caloric restriction. Moreover, mTOR signaling is reduced in the Ames Dwarf mouse, a model of extended longevity (Sharp and Bartke, 2005). Indeed, TOR inhibition extends lifespan in a variety of model organisms, and can even extend lifespan when inhibited during a limited period of adult life in mice (Harrison et al., 2009).

The mechanisms by which TOR exerts its effects on aging may involve the modulation of protein synthesis and autophagy (Medvedik et al., 2007; Steffen et al., 2008) (Chen et al., 2007; Hansen et al., 2007; Kapahi et al., 2004; Pan et al., 2007; Steffen et al., 2008). In mammals, mTOR functions in two distinct signaling complexes - mTORC1 and mTORC2. The mTORC1 complex is composed of mTOR, RAPTOR, PRAS40 and mLST8 and regulates cell growth, protein synthesis, ribosome biogenesis and autophagy through the activation of key downstream targets that include ribosomal S6 kinase and 4E-BP1 (Guertin and Sabatini, 2007). The mTORC2 complex is composed of mTOR, RICTOR, mSIN1, PROTOR and mLST8. Downregulation of mTORC1 activity by rapamycin reduces glycolysis and facilitates a switch in fat metabolism by increasing fatty acid oxidation in skeletal muscle cells (Brown et al., 2007; Sipula et al., 2006). S6K1−/− mice have reduced fat stores and are protected from weight gain on a high fat diet (Um et al., 2004), while mice lacking mTORC1 inhibitory targets 4E-BP1 and 4E-BP2 display diminished lipolysis, increased fatty acid synthesis and enhanced sensitivity to diet-induced obesity (Le Bacquer et al., 2007). TOR signaling therefore stands at the crossroads of metabolism, stress responses and aging, with potentially important implications for the pathogenesis and treatment of age-related metabolic disorders.

Mitochondrial Function, Oxidative Stress and Aging

Research spanning decades has investigated the idea that reactive oxygen species (ROS) generation leads to the macromolecular damage that underlies the aging process, known as the “Free Radical Theory of Aging” (Harmon, 1956). While many studies find a correlation between oxidative damage and lifespan, there are enough exceptions and contradictions to rule out simple causality (Lapointe and Hekimi, 2010). For example, in worms, a mutation of a subunit in complex II of the electron transport chain, succinate dehydrogenase cytochrome b, results in reduced respiration and decreased lifespan, especially under high oxygen conditions (Ishii et al., 1998), demonstrating a correlation between ROS and lifespan. In flies, decreasing superoxide dismutase (SOD) 2 levels in skeletal muscle decreases locomotion and lifespan (Martin et al., 2009), while increasing SOD2 promotes lifespan extension (Sun and Tower, 1999). By contrast, decreasing SOD isoforms in worms increases both ROS and lifespan (Van Raamsdonk and Hekimi, 2009). Manipulating MnSOD in mice also leads to mixed results; mice hetereozygous for MnSOD deletion demonstrate elevated ROS, but display normal lifespan, and MnSOD overexpression decreases lipid peroxidation and increases resistance against paraquot-induced oxidative stress, but does not extend lifespan (Jang et al., 2009). However, some mammalian studies support a link between ROS generation and lifespan. This was dramatically demonstrated by overexpression of catalase targeted specifically to mitochondria, which reduced oxidative stress and extended lifespan (Schriner et al., 2005). Catalase expression in the nucleus or peroxisomal compartments did not affect lifespan, emphasizing the central role of mitochondria in oxidative metabolism and lifespan regulation (Schriner et al., 2005).

Mice with a homozygous mutation in the exonuclease domain of mitochondria DNA (mtDNA) polymerase gamma (POLG) have been used as a model of mitochondrial dysfunction and aging. These mice possess a mtDNA mutator phenotype, accumulating large numbers of deletions and point mutations in mtDNA (Kujoth et al., 2005; Trifunovic et al., 2004). Surprisingly, these mice do not display signs of elevated ROS generation, but instead exhibit increased apoptosis, a number of age-related phenotypes, and a shortened lifespan (Kujoth et al., 2005; Trifunovic et al., 2004). Subsequent studies revealed that mtDNA deletions that accumulate in the brain and heart and provide the driving force behind the progeroid-like phenotype of the mutator mice (Vermulst et al., 2007; Vermulst et al., 2008). Interestingly, catalase overexpression can attenuate the age-dependent cardiomyopathy observed in POLG mutant mice (Dai et al., 2010), suggesting that oxidative stress may in fact contribute to some of the age-related phenotypes in this mouse model. Taken together, these studies demonstrate the critical but complex connection between mitochondrial function and lifespan.

Studies in model organisms show that mtDNA mutations can both reduce or extend lifespan, depending on severity, context and developmental stage. Surprisingly, complete absence of mtDNA in yeast, the so-called petite mutation, is associated with increased lifespan (Powell et al., 2000). Similarly in worms, RNA interference (RNAi) studies showed that decreasing the expression of mitochondrial genes increased lifespan (Dillin et al., 2002; Lee et al., 2003). Furthermore, timing studies showed that respiration must be decreased during development for the life-extending benefits (Dillin et al., 2002). Rea and colleagues demonstrated the importance of the level of mitochondrial gene expression in this effect (Rea et al., 2007). Moderate inhibition extended lifespan significantly, whereas high levels of RNAi inhibition reduced lifespan. The exact mechanisms that mediate lifespan extension in these models are still not known, but may involve a mitochondrial-driven stress response that is similar to hormesis. This idea is supported by glucose restriction in worms, an intervention that increases lifespan but is associated with increased mitochondrial respiration (Schulz et al., 2007). Remarkably, in this model, increased oxidative stress is required for lifespan benefits, as antioxidant treatment of glucose-restricted worms blocks lifespan extension (Schulz et al., 2007). Along these lines, feeding worms low doses of the ROS-generating compound juglone induces small heat shock protein HSP-16.2 and increases lifespan in a DAF-16-dependent manner. Administration of higher juglone concentrations results in decreased lifespan (Hartwig et al., 2009). Furthermore, in worms, SOD2 deletion and ubiquinone-defective clk-1mutants both show increased lifespan despite elevated ROS generation and oxidative stress (Lapointe and Hekimi, 2008; Van Raamsdonk and Hekimi, 2009). Taken together, these studies demonstrate that mild forms of mitochondrial dysfunction may activate stress response pathways that promote a protective environment, conducive to long life. A greater understanding of the mechanisms involved in mitochondrial stress response pathways might provide new therapeutic opportunities for aging and age-related pathology.

Mitochondrial function is controlled by a number of signaling pathways and transcriptional regulators that sense energetic stress and contribute to lifespan regulation, including peroxisome proliferation-activated receptor coactivator 1 α (PGC-1α), sirtuins, mTOR, and AMPK. When energy is low, these pathways allow the cell to adjust fuel utilization and mitochondrial number. Many aspects of mitochondrial dysfunction associated with aging can be blocked by caloric restriction. (Hunt et al., 2006; Sohal and Weindruch, 1996). For example, caloric restriction partially prevents the age-related decline in mitochondrial gene expression in mouse heart, brain and skeletal muscle (Lee et al., 2002; Lee et al., 1999; Lee et al., 2000). These metabolic effects are mediated by a complex interplay between signaling pathways that converge on the transcriptional co-activator PGC-1 α. PGC-1 α is a central regulator of mitochondrial biogenesis and function that is induced by a variety of metabolic stressors, including low energy availability and oxidative stress (Kelly and Scarpulla, 2004). Studies in mammalian cells and tissues have shown that caloric restriction induces mitochondrial biogenesis through the up-regulation of PGC-α (Lopez-Lluch et al., 2006) and endothelial nitric oxide synthase (eNOS) (Nisoli et al., 2005). Moreover, the severe muscle wasting phenotype of mice lacking a subunit of electron transport complex IV in skeletal muscle can be partially rescued by PGC-1 α overexpression, suggesting that boosting the total number of mitochondria may compensate for a mutation in the electron transport chain (Wenz et al., 2008). Likewise, age-associated sarcopenia and metabolic dysfunction can be rescued by PGC-1α overexpression (Wenz et al., 2009). SIRT1 also activates PGC-1α by deacetylation and promotes mitochondrial biogenesis (Mattagajasingh et al., 2007; Rodgers et al., 2005). Treatment with resveratrol results in increases the lifespan of mice fed a high fat diet (Baur et al., 2006; Feige et al., 2008), and this may be mediated, in part, via activation of PGC-1α by SIRT1. PGC-1α can also be activated by phosphorylation by AMPK, increasing the expression of target genes involved in mitochondrial biogenesis and fatty acid oxidation (Jager et al., 2007; Long et al., 2005). Hence, PGC-1α is a central node of regulation for several signaling pathways that regulate both mitochondrial function and lifespan.

Several studies have linked the TOR signaling pathway to altered mitochondrial function, nutrient sensing and lifespan regulation. In yeast, TOR inhibition regulates mitochondrial respiration and increases chronological lifespan (Bonawitz et al., 2007). Mice deficient for the TOR target S6K1 show increased expression of genes that mediate mitochondrial respiratory function and fatty acid oxidation in white adipose tissue and skeletal muscle (Um et al., 2004). Likewise, inhibition of mTOR signaling with rapamycin treatment decreases transcription of genes that augment mitochondrial function, such as PGC-1α, PGC-1β, NRF-1 and ERRα, resulting in reduced oxygen consumption and mitochondrial number. Under normal conditions, mTORC1 complexes with PGC-1α and the transcription factor YY1, leading to the transcription of nuclear-encoded mitochondrial genes. Upon treatment with rapamycin, the complex is dissociated downregulating the transcription of mitochondrial genes (Cunningham et al., 2007). These studies suggest that mTOR may link the control of lifespan to the regulation of cell growth and mitochondrial metabolism.

Genotoxic Stress and Aging

A distinct category of genetic disorders involving impaired sensing or repair of DNA damage has provided evidence for a central role of genome maintenance in the aging process. These disorders are known as segmental progerias because they are associated with a subset of the phenotypic changes that occur during normal aging, often including neurodegeneration and cancer. A prototypical example is Werner syndrome, which typically starts at puberty with accelerated onset of many different features of normal aging including cataracts, skin atrophy, hair loss, osteoporosis, atherosclerosis, type II diabetes and a variety of different neoplasms (Martin, 2005). Werner syndrome is caused by autosomal recessive mutations in a RecQ helicase that is involved in transcription, DNA replication and DNA repair (Rossi et al., 2010). In culture, cells from Werner syndrome patients exhibit genomic instability and accelerated senescence, features shared by other progeroid disorders caused by loss of function mutations in DNA repair genes, including Xeroderma pigmentosa, Cockayne syndrome, and trichothiodystrophy. Conversely, increased longevity may be associated with more efficient DNA repair. The lifespan regulating protein SIRT1 can promote DNA repair by deacetylation of repair proteins (Fan and Luo, 2010) and the double strand break sensor NBS1 (Yuan et al., 2007), and possibly by epigenetic modification of chromatin (Oberdoerffer et al., 2008).

The study of human progeroid syndromes has provided evidence for a link between genotoxic stress responses and signaling pathways that regulate the aging process. This was illustrated by a recent mouse model of a human progeroid syndrome caused by mutation of the XPF-ERCC1 endonuclease, which is involved in nucleotide excision repair. The transcriptional profile of the liver in XPF-ERCC1-deficient mice showed reduced expression of genes in the insulin/IGF-1 signaling pathway, and a shift to anabolic metabolism and increased anti-oxidant defense (Niedernhofer et al., 2006). Downregulation of insulin/IGF-1 signaling is also a feature of other mouse models with genomic instability, such as the SIRT6-knockout and a transgenic mouse with overexpression of a truncated p53 isoform (Maier et al., 2004; Mostoslavsky et al., 2006). An expression profile suggesting reduced insulin/IGF-1 signaling was also observed in cultured fibroblasts from Werner syndrome patients (Kyng and Bohr, 2005). The downregulation of this central signaling pathway may be a mechanism for shifting cellular resources from growth to repair and protection in the setting of genomic instability. Interestingly, caloric restriction or genetic mutations that extend lifespan have a similar effect on insulin/IGF-1 signaling in a variety of model organisms. Hence, genotoxic stress and aging may stimulate similar stress response pathways.

The broad spectrum of genotoxic stress responses and their potential relationship to aging is exemplified by the disorder ataxia telangiectasia (AT), which is caused by autosomal recessive mutations in the PI3 kinase-related kinase ataxia telangiectasia mutated (ATM). AT is characterized by immunodeficiency, skin lesions, pigmentary changes, neurodegeneration and a variety of neoplasms (Martin, 2005). The clinical manifestations of AT suggest that an impaired DNA damage response has different consequences in mitotic versus post-mitotic cells. Loss of efficient ATM-mediated signaling in mitotic cells leads to genomic instability, giving rise to immunodeficiency and a variety of tumors (Lombard et al., 2005). In post-mitotic neurons, however, the phenotype is one of degeneration, with a predilection for the large metabolically active Purkinje cell neurons of the cerebellum.

A recent proteomic analysis of the substrates of the ATM and ATR kinases suggests that DNA damage can activate a broad range of signaling pathways, some of which also regulate aging (Matsuoka et al., 2007). ATM is activated through autophosphorylation following interaction with the MRN complex (mre11-rad50-nbs1) at double strand breaks, whereas ATR is activated through interaction with ATR-interacting protein (ATRIP) at sites of stalled replication forks or single stranded DNA associated with double strand breaks. A proteome-wide analysis identified over 700 protein substrates of ATM and ATR that are phosphorylated in response to ionizing radiation (Matsuoka et al., 2007). In addition to replicating previously known DNA damage response proteins, many new substrates were identified including components of signaling pathways that regulate aging. For example, several proteins in the insulin/IGF-1 signaling pathway were identified, including IRS2, the kinase AKT3, and the transcription factor FOXO1. Multiple components of the protein translation regulatory pathway regulated by TOR signaling were also identified as ATM/ATR substrates, including TSC1, 4E-BP1 and p70S6K (ribosomal protein S6 kinase). Activation of these signaling pathways would be predicted to augment short-term cell survival by preventing apoptosis and increasing macromolecular biosynthesis. Surprisingly, an opposite effect, reduced insulin/IGF signaling, has been described in progeroid syndromes with chronic genomic instability. Hence, the DNA damage response network may have evolved to facilitate repair and survival in the short-term following acute DNA damage. However, repetitive or sustained activation of the DNA damage response during aging may compromise normal tissue homeostasis, leading to apoptosis or cellular senescence.

The notion that chronic activation of the DNA damage response may contribute to the aging process is supported by the phenotypes of different p53 gain of function mouse models. It was originally reported that a mutant p53 allele with enhanced tumor suppressor activity accelerated aging in a variety of different tissues in a mouse model (Tyner et al., 2002). However, accelerated aging was not observed in a different mouse model that overexpressed wild-type p53 under the control of the endogenous promoter (Garcia-Cao et al., 2002). A salient difference between the two models was the constitutive expression of the gain of function p53 allele versus the regulated expression of the allele under the control of the endogenous promoter, which was overexpressed under stress-related conditions, but not in the absence of stress. The constitutive overexpressing mutant mouse showed accelerated aging, whereas the stress-related expressor had a normal lifespan with augmented tumor suppressor activity. Hence, persistent stimulation of the DNA damage response resulting in chronic p53 activation, even at a low level, may be deleterious, potentially leading to apoptosis or cellular senescence (Lombard et al., 2005). Moreover, recent studies suggest that chronic DNA damage and ATM signaling in senescent cells leads to secretion of pro-inflammatory cytokines, possibly through activation of the ATM target NF-κB (Rodier et al., 2009). A persistent DNA damage response may therefore contribute to systemic inflammation, a known contributory factor for many age-related degenerative disorders.

A consequence of chronic DNA damage in the aging brain may be transcriptional repression and altered neuronal function (Lu et al., 2004). Transcriptional repression of actively expressed synaptic genes is pronounced in the aging human brain, especially for genes involved in cognitive and affective functions (Loerch et al., 2008). These genes may be selectively vulnerable to DNA damage owing to their activated euchromatic state in neurons, resulting in greater access of DNA to reactive oxygen species and other damaging agents. It is unclear, however, why DNA damage becomes persistent in aging neurons but is efficiently repaired in younger neurons. Another unresolved issue is the role of DNA damage in the pathogenesis of age-related neurodegenerative disorders, particularly Alzheimer’s and Parkinson’s disease, which are accompanied by substantial oxidative stress (Moreira et al., 2008).

Age-Dependent Decline in Proteostasis

A set of transcription factors, molecular chaperones and cofactors function together to promote correct protein folding and protect the cell by sequestering misfolded proteins in a process collectively known as proteostasis. The efficiency of this quality control system declines with age together with changes in protein structure due to oxidative modification, missense mutations and misincorporation of amino acids during translation. Compartments that are highly sensitive to redox state, such as the endoplasmic reticulum and mitochondria, are particularly vulnerable and have their own distinct protein folding quality control systems. The unfolded protein response (UPR) in the endoplasmic reticulum (ER) is activated by misfolding of newly synthesized proteins, and can be induced by exogenous toxins or by metabolic disorders such as type II diabetes (Ron and Walter, 2007). A similar but less well-defined pathway is induced by protein misfolding in mitochondria (Broadley and Hartl, 2008). In the cytoplasm and nucleus, a number of molecular chaperones, exemplified by hsp40, 70 and 90, monitor, sequester and promote the refolding of misfolded proteins. The cross-talk between protein quality control pathways across different cellular compartments is not well understood, although excessive protein misfolding in one compartment can globally affect proteostasis and may contribute to aging and the pathogenesis of neurodegenerative disorders (Bennett et al., 2005; Bennett et al., 2007; Kaganovich et al., 2008; Morimoto, 2008).

Genetic studies in C. elegans suggest that aging and proteostasis are closely linked and coordinately regulated. Aging worms show impaired activation of heat shock and unfolded protein responses and the accumulation of aggregated proteins (Ben-Zvi et al., 2009; David et al., 2010). Proteostasis in aging worms can be restored by manipulations that also extend lifespan, such as downregulation of insulin/IGF-1 signaling by RNAi or activation of the transcription factors DAF-16 and HSF-1 (David et al., 2010; Hsu et al., 2003; Morley et al., 2002). HSF-1 is a member of a family of transcription factors that act in worms and mammals to transcriptionally activate the proteostasis network. Lifespan extension from reduced insulin/IGF-1 signaling is suppressed by RNAi for HSF-1 or chaperones that are transcriptional targets of HSF-1, suggesting that proteostasis is a major component of lifespan regulation in the worm (Morley and Morimoto, 2004). In mammalian cells, the related transcription factor HSF is subject to feedback inhibition through interaction with shock proteins and by direct acetylation. Deacetylation of HSF by the stress resistance protein SIRT1 potentiates the transactivation of heat shock genes (Westerheide et al., 2009). Proteostasis is also a component or hormesis, in which mild stress results in a long-term increase in stress resistance. In worms, hormetic heat shock in young adult animals extends lifespan through the induction of heat shock proteins (Olsen et al., 2006).

Loss of function in the protein quality control network contributes to the pathogenesis of a group of age-dependent disorders known as conformational diseases that are exemplified by Alzheimer’s and Huntington’s disease. A characteristic feature is age-dependent accumulation of protein aggregates in the brain that may contribute to neurodegeneration and neurological decline. Huntington’s disease (HD) is the prototype of a group of human genetic diseases caused by polyglutamine-containing proteins that aggregate, leading to neurodegeneration. The length of the polyglutamine tract is inversely related to the age of onset of disease, suggesting a critical interplay between aging, proteostasis, and neurodegenerative pathology. Moreover, signaling pathways that regulate aging also control polyglutamine-related pathology in C. elegans models; loss of function mutants in the insulin/IGF-1 regulated transcription factors DAF-16 and HSF-1 reduce lifespan and accelerate polyglutamine aggregation and toxicity (Hsu et al., 2003). Conversely, the AGE-1 mutation that extends lifespan in worms reduces polyglutamine aggregation and toxicity (Morley et al., 2002). Conformational diseases also illustrate the fragility of the proteostasis network during aging. Misfolding and aggregation of a single metastable protein, such as mutant huntingtin, can impair the folding of other proteins (David et al., 2010; Gidalevitz et al., 2006) and globally inhibit the ubiquitin proteosome system (Bennett et al., 2007).

Recent studies in worms and mice suggest that stress response pathways which regulate proteostasis during aging may also contribute to the pathogenesis of Alzheimer’s disease (AD) (De Strooper, 2010). One of the first examples was a worm model in which the 42 amino acid form of the amyloid β-peptide (Aβ42), which forms deposits in the brain in AD, was expressed in muscle cells. This resulted in age-dependent aggregation and toxicity that could be prevented by downregulating the insulin/IGF-1 signaling pathway. The protective effect was mediated by activation of the transcription factors DAF-16 and HSF-1. Surprisingly, DAF-16 and HSF-1 had opposite effects on Aβ aggregation. HSF-1 activation resulted in the inhibition of Aβ aggregation, whereas DAF-16 promoted Aβ aggregation, but presumably through a more benign pathway that generated non-toxic fibrillar aggregates (Cohen et al., 2006). A significant modulating effect of insulin/IGF-1 signaling was also observed in APP transgenic mouse models, where downregulation of the pathway was also protective, reducing neuronal loss and inflammation, and improving cognitive function. This was paradoxically associated with increased Aβ aggregation in dense core amyloid plaques, similar to the protective effect of DAF-16 activation in the worm model. It was suggested that this unique proteostatic mechanism may protect against Aβ toxicity by inducing the formation of larger, more inert aggregates, resulting in less accumulation of smaller more toxic oligomeric forms (Cohen et al., 2006). Hence, age-related neurodegeneration may be modulated by insulin/IGF-1 signaling through the control of proteostasis and other stress response pathways.

Autophagy in aging

Autophagy declines in nearly all cells and tissue types as organisms age, likely contributing to the accumulation of dysfunctional organelles and damaged proteins (Cuervo et al., 2010). Studies using model organisms have firmly established a direct role for autophagy in the regulation of lifespan extension. A recent genome-wide screen in yeast identified autophagy genes as a requirement for normal chronological lifespan (Fabrizio et al., 2010). In worms, increased autophagy alone is not sufficient to promote lifespan extension, but is required for the increased lifespan extension due to calorie restriction or decreased insulin signaling (Hansen et al., 2008). In flies, the expression of autophagy genes declines in neurons, and increasing autophagy in these cells promotes lifespan extension, while reducing autophagy in neurons decreases lifespan, increases oxidative stress and results in neurodegeneration (Juhasz et al., 2007; Simonsen et al., 2008). Feeding rapamycin to adult flies increases stress resistance and extends their lifespan via alterations in both autophagy and translation (Bjedov et al., 2010).

Mouse studies demonstrate the central importance of autophagy in mammalian aging and age-related neurodegeneration. Mice deficient for autophagy-related genes (ATGs) exhibit reduced autophagy, high levels of protein inclusion bodies, and neurodegenerative pathology (Hara et al., 2006; Komatsu et al., 2006). These phenotypes are also common features of human neurodegenerative diseases associated with protein aggregation, such as Alzheimer’s disease and Huntington’s disease (Bishop et al., 2010; Cuervo et al., 2010), suggesting a close relationship between the regulation of autophagy, aging and neurodegeneration. Moreover, blocking mTOR in a mouse model of AD decreases Aβ levels and improves cognitive function, and these effects may be due, in part, to increased autophagy (Spilman et al., 2010). Thus, the lifespan extension observed in mammals treated with mTOR inhibitor rapamycin (Harrison et al., 2009) may also be mediated, in part, by the up-regulation of autophagy. Taken together, these studies suggest a role for autophagy in the coordinated regulation of aging and age-related stress responses, as well as the pathogenesis of neurodegenerative disorders.

Activation of autophagy is fundamentally linked to the nutrient status of the cell and impacts metabolism and stress response pathways. Autophagy increases during times of nutrient deprivation, such as in fasting or calorie restriction. This is likely to provide cells with an additional source of energy during times of nutrient scarcity; generating free amino acids can be utilized for protein synthesis. Autophagy is also important for mobilizing lipids in the liver during nutrient deprivation (Singh et al., 2009a), which can be metabolized as an alternate energy source, or used for new membrane biosynthesis. Strikingly, overexpression of LAMP-2A in mouse liver in the late stages of adult life prevents an age-related decline in autophagy, reducing the associated accumulation of damaged proteins and leading to improved liver function (Zhang and Cuervo, 2008).

The mammalian sirtuin SIRT1 stimulates rapamycin- and starvation-stimulated autophagy in cultured cells via deacetylation and activation of autophagy proteins Atg5, Atg7 and Atg8 (Lee et al., 2008), suggesting that at least some of the effects of SIRT1 on metabolism and stress responses could be due to regulation of this clearance pathway. SIRT1 was also required for the CR-mediated increase in mitochondrial autophagy in mouse kidney observed during hypoxia (Kume et al., 2010), demonstrating multiple levels by which protein turnover can be regulated by this sirtuin. Hence, autophagy may be a major component of age-dependent stress response pathways that are activated by central regulators such as SIRT1 and mTOR.

The necessity for post-mitotic neurons to survive an entire lifespan imposes a major challenge for protein quality control, in particular autophagic clearance of damaged or aggregated proteins. Age-related diseases of the nervous system are characterized by the accumulation of aggregated proteins in affected neurons and glial cells that may contribute to dysfunction and neurodegeneration. Autophagy can become impaired in the aging brain at a number of levels. The initial formation of the autophagosome is mediated by the Class III PI3 kinase complex composed of beclin-1, Vps15 and Vps34. Beclin-1 levels decline in the normal aging brain (Shibata et al., 2006) and fall further in early AD, potentially reducing autophagy and predisposing to Aβ aggregation (Pickford et al., 2008). Autophagosome formation may also be impaired at the level of cargo recognition. This has been observed in models of Huntington’s disease, resulting in impaired clearance of aggregated proteins despite normal autophagosome formation (Martinez-Vicente et al., 2010). There is also evidence of impaired autophagosome clearance in neurodegenerative disorders, resulting in the accumulation of autophagosomes that are not efficiently processed further (Wong and Cuervo, 2010). This may relate, in part, to disruption of microtubular or actin cytoskeletal elements that are required for autophagosome transport and fusion. Autophagosome accumulation may contribute directly to neurodegeneration by serving as a source of cytotoxic proteins, such as the 42 amino acid form of Aβ, or by disrupting cellular trafficking (Yu et al., 2005). Presenilin-1 mutations that cause familial AD have recently been shown to compromise both delivery to the lysosome and acidification of lysosomes, compromising the clearance of autophagic substrates at several levels (Lee et al., 2010). Deficits in autophagy can also occur in specific subcellular compartments. Two genes associated with familial Parkinson’s disease, PINK1 and parkin, may function together to promote autophagy of failing mitochondria, a process known as mitophagy (Geisler et al., 2010; Narendra et al., 2008). Disease-causing mutations in PINK1 and parkin impair this process, potentially leading to the persistence of dysfunctional mitochondria that generate high levels of reactive oxygen species and increase oxidative stress. These observations provide evidence for a broad range of mechanisms leading to impaired autophagy in age-related neurodegenerative disorders.

The complex role of autophagy in aging and age-related diseases is apparent in the varying outcomes and context-dependence of manipulating the autophagic pathway. Although activating autophagy is often beneficial, inhibiting autophagosome formation may also be beneficial by reducing neurodegeneration when autophagosome clearance is impaired, as in frontotemporal dementia (Lee and Gao, 2009). Another example is inhibition of the autophagy proteins ATG-5 or ATG-7 in white adipose tissue in the developing mouse. This decreases adipocyte differentiation, resulting in a lean mouse with increased insulin sensitivity. It also confers some of the positive metabolic features of brown adipocytes, such as increased fatty acid oxidation (Singh et al., 2009b). Hence, pharmacologic manipulation of the autophagic pathway has great therapeutic potential, but will likely need to be targeted to specific cell types in defined contexts.

Conclusion

The insulin/IGF, TOR and sirtuin networks integrate cellular and organismal homeostasis through coordinate regulation of nutrient sensing and stress response pathways. These signaling pathways respond to changes in nutrient status and various stressors by altering mitochondrial and metabolic function, and mobilizing the genome maintenance and proteostasis networks. The integrated action of these stress response and maintenance systems has been optimized in the early years of life to maximize fitness. Many studies suggest, however, that a decline in the effectiveness and integration of stress responses contributes to aging and age-related diseases. One of the more remarkable insights from aging research during the past two decades is that age-related decline is not invariably fixed, but can be modified by augmenting stress resistance through the conserved signaling pathways, leading to lifespan extension. A greater understanding of the systems biology of stress response signaling and its breakdown during aging may lead to new therapeutic approaches to the intractable degenerative disorders of aging.

Figure 1.

Figure 1

Age-related stress and disease. Aging is associated with mitochondrial dysfunction leading to reduced respiratory metabolism and increased generation of reactive oxygen species (ROS). Persistent DNA damage may arise from both increased oxidative damage and reduced efficiency of energy-intensive DNA repair, predisposing to apoptosis, senescence and inflammation. Aging is also associated with increased protein misfolding and aggregation in the cytoplasm, nucleus and endoplasmic reticulum. The various sites of age-related cellular damage and the physiological decline that ensues contribute to the pathogenesis of age-related diseases, including metabolic syndrome, inflammatory disorders, cancer, and neurodegenerative diseases.

Figure 2.

Figure 2

Nutritional regulation of conserved signaling and stress response pathways. Dietary restriction extends lifespan and augments stress resistance in many species by altering cellular metabolism and mobilizing protective stress responses. Gene and protein networks that maintain mitochondrial function, genomic stability and proteostasis are coordinately regulated by insulin/IGF and TOR signaling, and modulated by sirtuins.

Acknowledgements

We apologize for many studies that were not discussed because of limited space. This work is supported by grants from the National Institutes of Health/National Institute on Aging (AG26651, AG27916 and AG036106) to B.A.Y. and (AG032375) to M.C.H, a New Scholar Award from the Ellison Medical Foundation to M.C.H., and funding from the Glenn Foundation for Medical Research to B.A.Y. and M.C.H.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributor Information

Marcia C. Haigis, Email: marcia_haigis@hms.harvard.edu.

Bruce A. Yankner, Email: bruce_yankner@hms.harvard.edu.

References

  1. Abu-Elheiga L, Matzuk MM, Abo-Hashema KA, Wakil SJ. Continuous fatty acid oxidation and reduced fat storage in mice lacking acetyl-CoA carboxylase 2. Science. 2001;291:2613–2616. doi: 10.1126/science.1056843. [DOI] [PubMed] [Google Scholar]
  2. Banks AS, Kon N, Knight C, Matsumoto M, Gutierrez-Juarez R, Rossetti L, Gu W, Accili D. SirT1 gain of function increases energy efficiency and prevents diabetes in mice. Cell Metab. 2008;8:333–341. doi: 10.1016/j.cmet.2008.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Baur JA, Pearson KJ, Price NL, Jamieson HA, Lerin C, Kalra A, Prabhu VV, Allard JS, Lopez-Lluch G, Lewis K, et al. Resveratrol improves health and survival of mice on a high-calorie diet. Nature. 2006;444:337–342. doi: 10.1038/nature05354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Ben-Zvi A, Miller EA, Morimoto RI. Collapse of proteostasis represents an early molecular event in Caenorhabditis elegans aging. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:14914–14919. doi: 10.1073/pnas.0902882106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bennett EJ, Bence NF, Jayakumar R, Kopito RR. Global impairment of the ubiquitin-proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Molecular Cell. 2005;17:351–365. doi: 10.1016/j.molcel.2004.12.021. [DOI] [PubMed] [Google Scholar]
  6. Bennett EJ, Shaler TA, Woodman B, Ryu KY, Zaitseva TS, Becker CH, Bates GP, Schulman H, Kopito RR. Global changes to the ubiquitin system in Huntington's disease. Nature. 2007;448:704–708. doi: 10.1038/nature06022. [DOI] [PubMed] [Google Scholar]
  7. Bishop NA, Guarente L. Two neurons mediate diet-restriction-induced longevity in C. elegans. Nature. 2007;447:545–549. doi: 10.1038/nature05904. [DOI] [PubMed] [Google Scholar]
  8. Bishop NA, Lu T, Yankner BA. Neural mechanisms of ageing and cognitive decline. Nature. 2010;464:529–535. doi: 10.1038/nature08983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bjedov I, Toivonen JM, Kerr F, Slack C, Jacobson J, Foley A, Partridge L. Mechanisms of life span extension by rapamycin in the fruit fly Drosophila melanogaster. Cell Metab. 2010;11:35–46. doi: 10.1016/j.cmet.2009.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Boily G, Seifert EL, Bevilacqua L, He XH, Sabourin G, Estey C, Moffat C, Crawford S, Saliba S, Jardine K, et al. SirT1 regulates energy metabolism and response to caloric restriction in mice. PLoS ONE. 2008;3:e1759. doi: 10.1371/journal.pone.0001759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bonawitz ND, Chatenay-Lapointe M, Pan Y, Shadel GS. Reduced TOR signaling extends chronological life span via increased respiration and upregulation of mitochondrial gene expression. Cell Metab. 2007;5:265–277. doi: 10.1016/j.cmet.2007.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bonkowski MS, Dominici FP, Arum O, Rocha JS, Al Regaiey KA, Westbrook R, Spong A, Panici J, Masternak MM, Kopchick JJ, Bartke A. Disruption of growth hormone receptor prevents calorie restriction from improving insulin action and longevity. PloS one. 2009;4:e4567. doi: 10.1371/journal.pone.0004567. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  13. Bordone L, Cohen D, Robinson A, Motta MC, van Veen E, Czopik A, Steele AD, Crowe H, Marmor S, Luo J, et al. SIRT1 transgenic mice show phenotypes resembling calorie restriction. Aging Cell. 2007;6:759–767. doi: 10.1111/j.1474-9726.2007.00335.x. [DOI] [PubMed] [Google Scholar]
  14. Bordone L, Guarente L. Calorie restriction, SIRT1 and metabolism: understanding longevity. Nat Rev Mol Cell Biol. 2005;6:298–305. doi: 10.1038/nrm1616. [DOI] [PubMed] [Google Scholar]
  15. Broadley SA, Hartl FU. Mitochondrial stress signaling: a pathway unfolds. Trends in Cell Biology. 2008;18:1–4. doi: 10.1016/j.tcb.2007.11.003. [DOI] [PubMed] [Google Scholar]
  16. Brown NF, Stefanovic-Racic M, Sipula IJ, Perdomo G. The mammalian target of rapamycin regulates lipid metabolism in primary cultures of rat hepatocytes. Metabolism. 2007;56:1500–1507. doi: 10.1016/j.metabol.2007.06.016. [DOI] [PubMed] [Google Scholar]
  17. Buchberger A, Bukau B, Sommer T. Protein quality control in the cytosol and the Endoplasmic Reticulum: Brothers in Arms. Mol Cell. 2010;40:XXX. doi: 10.1016/j.molcel.2010.10.001. [DOI] [PubMed] [Google Scholar]
  18. Chen D, Pan KZ, Palter JE, Kapahi P. Longevity determined by developmental arrest genes in Caenorhabditis elegans. Aging Cell. 2007;6:525–533. doi: 10.1111/j.1474-9726.2007.00305.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Ciccia A, Elledge SJ. The DNA Damage Response: Making it safe to play with knives. Mol Cell. 2010;40:XXX. doi: 10.1016/j.molcel.2010.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Cohen E, Bieschke J, Perciavalle RM, Kelly JW, Dillin A. Opposing activities protect against age-onset proteotoxicity. Science. 2006;313:1604–1610. doi: 10.1126/science.1124646. [DOI] [PubMed] [Google Scholar]
  21. Comfort A. Effect of Delayed and Resumed Growth on the Longevity of a Fish (Lebistes Reticulatus, Peters) in Captivity. Gerontologia. 1963;49:150–155. doi: 10.1159/000211216. [DOI] [PubMed] [Google Scholar]
  22. Cuervo AM, Wong ES, Martinez-Vicente M. Protein degradation, aggregation, and misfolding. Mov Disord. 2010;25 Suppl 1:S49–S54. doi: 10.1002/mds.22718. [DOI] [PubMed] [Google Scholar]
  23. Cunningham JT, Rodgers JT, Arlow DH, Vazquez F, Mootha VK, Puigserver P. mTOR controls mitochondrial oxidative function through a YY1-PGC-1alpha transcriptional complex. Nature. 2007;450:736–740. doi: 10.1038/nature06322. [DOI] [PubMed] [Google Scholar]
  24. Dai DF, Chen T, Wanagat J, Laflamme M, Marcinek DJ, Emond MJ, Ngo CP, Prolla TA, Rabinovitch PS. Age-dependent cardiomyopathy in mitochondrial mutator mice is attenuated by overexpression of catalase targeted to mitochondria. Aging cell. 2010;9:536–544. doi: 10.1111/j.1474-9726.2010.00581.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. David DC, Ollikainen N, Trinidad JC, Cary MP, Burlingame AL, Kenyon C. Widespread Protein Aggregation as an Inherent Part of Aging in C. elegans. PLoS Biology. 2010;8 doi: 10.1371/journal.pbio.1000450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. De Strooper B. Proteases and proteolysis in Alzheimer disease: a multifactorial view on the disease process. Physiological reviews. 2010;90:465–494. doi: 10.1152/physrev.00023.2009. [DOI] [PubMed] [Google Scholar]
  27. Dillin A, Hsu AL, Arantes-Oliveira N, Lehrer-Graiwer J, Hsin H, Fraser AG, Kamath RS, Ahringer J, Kenyon C. Rates of behavior and aging specified by mitochondrial function during development. Science. 2002;298:2398–2401. doi: 10.1126/science.1077780. [DOI] [PubMed] [Google Scholar]
  28. Fabrizio P, Hoon S, Shamalnasab M, Galbani A, Wei M, Giaever G, Nislow C, Longo VD. Genome-wide screen in Saccharomyces cerevisiae identifies vacuolar protein sorting, autophagy, biosynthetic, and tRNA methylation genes involved in life span regulation. PLoS genetics. 2010;6:e1001024. doi: 10.1371/journal.pgen.1001024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Fan W, Luo J. SIRT1 regulates UV-induced DNA repair through deacetylating XPA. Molecular Cell. 2010;39:247–258. doi: 10.1016/j.molcel.2010.07.006. [DOI] [PubMed] [Google Scholar]
  30. Feige JN, Lagouge M, Canto C, Strehle A, Houten SM, Milne JC, Lambert PD, Mataki C, Elliott PJ, Auwerx J. Specific SIRT1 activation mimics low energy levels and protects against diet-induced metabolic disorders by enhancing fat oxidation. Cell Metab. 2008;8:347–358. doi: 10.1016/j.cmet.2008.08.017. [DOI] [PubMed] [Google Scholar]
  31. Flachsbart F, Caliebe A, Kleindorp R, Blanché H, von Eller-Eberstein H, Nikolaus S, Schreiber S, Nebel A. Association of FOXO3A variation with human longevity confirmed in German centenarians. Proc Natl Acad Sci U S A. 2009;106:2700–2705. doi: 10.1073/pnas.0809594106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Frye RA. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun. 2000;273:793–798. doi: 10.1006/bbrc.2000.3000. [DOI] [PubMed] [Google Scholar]
  33. Garcia-Cao I, Garcia-Cao M, Martin-Caballero J, Criado LM, Klatt P, Flores JM, Weill JC, Blasco MA, Serrano M. “Super p53” mice exhibit enhanced DNA damage response, are tumor resistant and age normally. EMBO Journal. 2002;21:6225–6235. doi: 10.1093/emboj/cdf595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, Kahle PJ, Springer W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol. 2010;12:119–131. doi: 10.1038/ncb2012. [DOI] [PubMed] [Google Scholar]
  35. Giannakou ME, Goss M, Partridge L. Role of dFOXO in lifespan extension by dietary restriction in Drosophila melanogaster: not required, but its activity modulates the response. Aging cell. 2008;7:187–198. doi: 10.1111/j.1474-9726.2007.00362.x. [DOI] [PubMed] [Google Scholar]
  36. Gidalevitz T, Ben-Zvi A, Ho KH, Brignull HR, Morimoto RI. Progressive disruption of cellular protein folding in models of polyglutamine diseases. Science. 2006;311:1471–1474. doi: 10.1126/science.1124514. [DOI] [PubMed] [Google Scholar]
  37. Greer EL, Brunet A. Different dietary restriction regimens extend lifespan by both independent and overlapping genetic pathways in C. elegans. Aging cell. 2009;8:113–127. doi: 10.1111/j.1474-9726.2009.00459.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Guarente L. Sirtuins as potential targets for metabolic syndrome. Nature. 2006;444:868–874. doi: 10.1038/nature05486. [DOI] [PubMed] [Google Scholar]
  39. Guertin DA, Sabatini DM. Defining the role of mTOR in cancer. Cancer Cell. 2007;12:9–22. doi: 10.1016/j.ccr.2007.05.008. [DOI] [PubMed] [Google Scholar]
  40. Haigis MC, Sinclair DA. Mammalian sirtuins: biological insights and disease relevance. Annual review of pathology. 2010;5:253–295. doi: 10.1146/annurev.pathol.4.110807.092250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Hansen M, Chandra A, Mitic LL, Onken B, Driscoll M, Kenyon C. A role for autophagy in the extension of lifespan by dietary restriction in C. elegans. PLoS genetics. 2008;4:e24. doi: 10.1371/journal.pgen.0040024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hansen M, Taubert S, Crawford D, Libina N, Lee SJ, Kenyon C. Lifespan extension by conditions that inhibit translation in Caenorhabditis elegans. Aging Cell. 2007;6:95–110. doi: 10.1111/j.1474-9726.2006.00267.x. [DOI] [PubMed] [Google Scholar]
  43. Hara T, Nakamura K, Matsui M, Yamamoto A, Nakahara Y, Suzuki-Migishima R, Yokoyama M, Mishima K, Saito I, Okano H, Mizushima N. Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature. 2006;441:885–889. doi: 10.1038/nature04724. [DOI] [PubMed] [Google Scholar]
  44. Harmon D. Aging: a theory based on free radical and radiation chemistry. J. Gerontol. 1956;11:298. doi: 10.1093/geronj/11.3.298. [DOI] [PubMed] [Google Scholar]
  45. Harrison DE, Strong R, Sharp ZD, Nelson JF, Astle CM, Flurkey K, Nadon NL, Wilkinson JE, Frenkel K, Carter CS, et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature. 2009;460:392–395. doi: 10.1038/nature08221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Hartwig K, Heidler T, Moch J, Daniel H, Wenzel U. Feeding a ROS-generator to Caenorhabditis elegans leads to increased expression of small heat shock protein HSP-16.2 and hormesis. Genes & nutrition. 2009;4:59–67. doi: 10.1007/s12263-009-0113-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Herranz D, Munoz-Martin M, Canamero M, Mulero F, Martinez-Pastor B, Fernandez-Capetillo O, Serrano M. Sirt1 improves healthy ageing and protects from metabolic syndrome-associated cancer. Nature Communications. 2010;1:1–8. doi: 10.1038/ncomms1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hsu AL, Murphy CT, Kenyon C. Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science. 2003;300:1142–1145. doi: 10.1126/science.1083701. [DOI] [PubMed] [Google Scholar]
  49. Hunt ND, Hyun DH, Allard JS, Minor RK, Mattson MP, Ingram DK, de Cabo R. Bioenergetics of aging and calorie restriction. Ageing research reviews. 2006;5:125–143. doi: 10.1016/j.arr.2006.03.006. [DOI] [PubMed] [Google Scholar]
  50. Ishii N, Fujii M, Hartman PS, Tsuda M, Yasuda K, Senoo-Matsuda N, Yanase S, Ayusawa D, Suzuki K. A mutation in succinate dehydrogenase cytochrome b causes oxidative stress and ageing in nematodes. Nature. 1998;394:694–697. doi: 10.1038/29331. [DOI] [PubMed] [Google Scholar]
  51. Jager S, Handschin C, St-Pierre J, Spiegelman BM. AMP-activated protein kinase (AMPK) action in skeletal muscle via direct phosphorylation of PGC-1alpha. Proc Natl Acad Sci U S A. 2007;104:12017–12022. doi: 10.1073/pnas.0705070104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Jang YC, Perez VI, Song W, Lustgarten MS, Salmon AB, Mele J, Qi W, Liu Y, Liang H, Chaudhuri A, et al. Overexpression of Mn superoxide dismutase does not increase life span in mice. The journals of gerontology. 2009;64:1114–1125. doi: 10.1093/gerona/glp100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Juhasz G, Erdi B, Sass M, Neufeld TP. Atg7-dependent autophagy promotes neuronal health, stress tolerance, and longevity but is dispensable for metamorphosis in Drosophila. Genes and Development. 2007;21:3061–3066. doi: 10.1101/gad.1600707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Kaeberlein M, McVey M, Guarente L. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev. 1999;13:2570–2580. doi: 10.1101/gad.13.19.2570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Kaganovich D, Kopito R, Frydman J. Misfolded proteins partition between two distinct quality control compartments. Nature. 2008;454:1088–1095. doi: 10.1038/nature07195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Kahn BB, Alquier T, Carling D, Hardie DG. AMP-activated protein kinase: ancient energy gauge provides clues to modern understanding of metabolism. Cell Metab. 2005;1:15–25. doi: 10.1016/j.cmet.2004.12.003. [DOI] [PubMed] [Google Scholar]
  57. Kapahi P, Zid BM, Harper T, Koslover D, Sapin V, Benzer S. Regulation of lifespan in Drosophila by modulation of genes in the TOR signaling pathway. Curr Biol. 2004;14:885–890. doi: 10.1016/j.cub.2004.03.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Kelly DP, Scarpulla RC. Transcriptional regulatory circuits controlling mitochondrial biogenesis and function. Genes Dev. 2004;18:357–368. doi: 10.1101/gad.1177604. [DOI] [PubMed] [Google Scholar]
  59. Kenyon C, Chang J, Gensch E, Rudner A, Tabtiang R. A C. elegans mutant that lives twice as long as wild type. Nature. 1993;366:461–464. doi: 10.1038/366461a0. [DOI] [PubMed] [Google Scholar]
  60. Kenyon CJ. The genetics of ageing. Nature. 2010;464:504–512. doi: 10.1038/nature08980. [DOI] [PubMed] [Google Scholar]
  61. Kimura KD, Tissenbaum HA, Liu Y, Ruvkun G. daf-2, an insulin receptor-like gene that regulates longevity and diapause in Caenorhabditis elegans. Science. 1997;277:942–946. doi: 10.1126/science.277.5328.942. [DOI] [PubMed] [Google Scholar]
  62. Komatsu M, Waguri S, Chiba T, Murata S, Iwata J, Tanida I, Ueno T, Koike M, Uchiyama Y, Kominami E, Tanaka K. Loss of autophagy in the central nervous system causes neurodegeneration in mice. Nature. 2006;441:880–884. doi: 10.1038/nature04723. [DOI] [PubMed] [Google Scholar]
  63. Kroemer G, Marino G, Levine B. Autophagy and the integrated stress response. Mol Cell. 2010;40:XXX. doi: 10.1016/j.molcel.2010.09.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kujoth GC, Hiona A, Pugh TD, Someya S, Panzer K, Wohlgemuth SE, Hofer T, Seo AY, Sullivan R, Jobling WA, et al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science. 2005;309:481–484. doi: 10.1126/science.1112125. [DOI] [PubMed] [Google Scholar]
  65. Kume S, Uzu T, Horiike K, Chin-Kanasaki M, Isshiki K, Araki S, Sugimoto T, Haneda M, Kashiwagi A, Koya D. Calorie restriction enhances cell adaptation to hypoxia through Sirt1-dependent mitochondrial autophagy in mouse aged kidney. J Clin Invest. 2010;120:1043–1055. doi: 10.1172/JCI41376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Kyng KJ, Bohr VA. Gene expression and DNA repair in progeroid syndromes and human aging. Ageing research reviews. 2005;4:579–602. doi: 10.1016/j.arr.2005.06.008. [DOI] [PubMed] [Google Scholar]
  67. Lagouge M, Argmann C, Gerhart-Hines Z, Meziane H, Lerin C, Daussin F, Messadeq N, Milne J, Lambert P, Elliott P, et al. Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC-1alpha. Cell. 2006;127:1109–1122. doi: 10.1016/j.cell.2006.11.013. [DOI] [PubMed] [Google Scholar]
  68. Lapointe J, Hekimi S. Early mitochondrial dysfunction in long-lived Mclk1+/−mice. J Biol Chem. 2008;283:26217–26227. doi: 10.1074/jbc.M803287200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Lapointe J, Hekimi S. When a theory of aging ages badly. Cell Mol Life Sci. 2010;67:1–8. doi: 10.1007/s00018-009-0138-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Le Bacquer O, Petroulakis E, Paglialunga S, Poulin F, Richard D, Cianflone K, Sonenberg N. Elevated sensitivity to diet-induced obesity and insulin resistance in mice lacking 4E–BP1 and 4E–BP2. J Clin Invest. 2007;117:387–396. doi: 10.1172/JCI29528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Lee CK, Allison DB, Brand J, Weindruch R, Prolla TA. Transcriptional profiles associated with aging and middle age-onset caloric restriction in mouse hearts. Proc Natl Acad Sci U S A. 2002;99:14988–14993. doi: 10.1073/pnas.232308999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Lee CK, Klopp RG, Weindruch R, Prolla TA. Gene expression profile of aging and its retardation by caloric restriction. Science. 1999;285:1390–1393. doi: 10.1126/science.285.5432.1390. [DOI] [PubMed] [Google Scholar]
  73. Lee CK, Weindruch R, Prolla TA. Gene-expression profile of the ageing brain in mice. Nature genetics. 2000;25:294–297. doi: 10.1038/77046. [DOI] [PubMed] [Google Scholar]
  74. Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns NE, Tsokos M, Alt FW, Finkel T. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy. Proc Natl Acad Sci U S A. 2008;105:3374–3379. doi: 10.1073/pnas.0712145105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Lee JA, Gao FB. Inhibition of autophagy induction delays neuronal cell loss caused by dysfunctional ESCRT-III in frontotemporal dementia. Journal of Neuroscience. 2009;29:8506–8511. doi: 10.1523/JNEUROSCI.0924-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Lee JH, Yu WH, Kumar A, Lee S, Mohan PS, Peterhoff CM, Wolfe DM, Martinez-Vicente M, Massey AC, Sovak G, et al. Lysosomal proteolysis and autophagy require presenilin 1 and are disrupted by Alzheimer-related PS1 mutations. Cell. 2010;141:1146–1158. doi: 10.1016/j.cell.2010.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Lee SS, Lee RY, Fraser AG, Kamath RS, Ahringer J, Ruvkun G. A systematic RNAi screen identifies a critical role for mitochondria in C. elegans longevity. Nature genetics. 2003;33:40–48. doi: 10.1038/ng1056. [DOI] [PubMed] [Google Scholar]
  78. Lin SJ, Defossez PA, Guarente L. Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science. 2000;289:2126–2128. doi: 10.1126/science.289.5487.2126. [DOI] [PubMed] [Google Scholar]
  79. Loerch PM, Lu T, Dakin KA, Vann JM, Isaacs A, Geula C, Wang J, Pan Y, Gabuzda DH, Li C, Prolla TA, Yankner BA. Evolution of the aging brain transcriptome and synaptic regulation. PLoS ONE. 2008;3:e3329. doi: 10.1371/journal.pone.0003329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Lombard DB, Chua KF, Mostoslavsky R, Franco S, Gostissa M, Alt FW. DNA repair, genome stability, and aging. Cell. 2005;120:497–512. doi: 10.1016/j.cell.2005.01.028. [DOI] [PubMed] [Google Scholar]
  81. Long YC, Barnes BR, Mahlapuu M, Steiler TL, Martinsson S, Leng Y, Wallberg-Henriksson H, Andersson L, Zierath JR. Role of AMP-activated protein kinase in the coordinated expression of genes controlling glucose and lipid metabolism in mouse white skeletal muscle. Diabetologia. 2005;48:2354–2364. doi: 10.1007/s00125-005-1962-5. [DOI] [PubMed] [Google Scholar]
  82. Lopez-Lluch G, Hunt N, Jones B, Zhu M, Jamieson H, Hilmer S, Cascajo MV, Allard J, Ingram DK, Navas P, de Cabo R. Calorie restriction induces mitochondrial biogenesis and bioenergetic efficiency. Proc Natl Acad Sci U S A. 2006;103:1768–1773. doi: 10.1073/pnas.0510452103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Lu T, Pan Y, Kao SY, Li C, Kohane I, Chan J, Yankner BA. Gene regulation and DNA damage in the ageing brain. Nature. 2004;429:883–891. doi: 10.1038/nature02661. [DOI] [PubMed] [Google Scholar]
  84. Maier B, Gluba W, Bernier B, Turner T, Mohammad K, Guise T, Sutherland A, Thorner M, Scrable H. Modulation of mammalian life span by the short isoform of p53. Genes and Development. 2004;18:306–319. doi: 10.1101/gad.1162404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Majmundar AJ, Wong WJ, Simon CM. Hypoxia inducible factors and the response to hypoxic stress. Mol. Cell. 2010;40 doi: 10.1016/j.molcel.2010.09.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. 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]
  87. Martin I, Jones MA, Rhodenizer D, Zheng J, Warrick JM, Seroude L, Grotewiel M. Sod2 knockdown in the musculature has whole-organism consequences in Drosophila. Free radical biology & medicine. 2009;47:803–813. doi: 10.1016/j.freeradbiomed.2009.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Martinez-Vicente M, Talloczy Z, Wong E, Tang G, Koga H, Kaushik S, de Vries R, Arias E, Harris S, Sulzer D, Cuervo AM. Cargo recognition failure is responsible for inefficient autophagy in Huntington's disease. Nature Neuroscience. 2010;13:567–576. doi: 10.1038/nn.2528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, 3rd, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y, et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science. 2007;316:1160–1166. doi: 10.1126/science.1140321. [DOI] [PubMed] [Google Scholar]
  90. Mattagajasingh I, Kim CS, Naqvi A, Yamamori T, Hoffman TA, Jung SB, DeRicco J, Kasuno K, Irani K. SIRT1 promotes endothelium-dependent vascular relaxation by activating endothelial nitric oxide synthase. Proc Natl Acad Sci U S A. 2007;104:14855–14860. doi: 10.1073/pnas.0704329104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. McCarty MF. Chronic activation of AMP-activated kinase as a strategy for slowing aging. Med Hypotheses. 2004;63:334–339. doi: 10.1016/j.mehy.2004.01.043. [DOI] [PubMed] [Google Scholar]
  92. 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. discussion 172. [PubMed] [Google Scholar]
  93. Medvedik O, Lamming DW, Kim KD, Sinclair DA. MSN2 and MSN4 link calorie restriction and TOR to sirtuin-mediated lifespan extension in Saccharomyces cerevisiae. PLoS Biol. 2007;5:e261. doi: 10.1371/journal.pbio.0050261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Moreira PI, Nuomura A, Nakamura M, Takeda A, Shenk JC, Aliev G, Smith MA, Perry G. Nucleic acid oxidation in Alzheimer disease. Free Radic Biol and Med. 2008;44:1493–1505. doi: 10.1016/j.freeradbiomed.2008.01.002. [DOI] [PubMed] [Google Scholar]
  95. Morimoto RI. Proteotoxic stress and inducible chaperone networks in neurodegenerative disease and aging. Genes and Development. 2008;22:1427–1438. doi: 10.1101/gad.1657108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Morley JF, Brignull HR, Weyers JJ, Morimoto RI. The threshold for polyglutamine-expansion protein aggregation and cellular toxicity is dynamic and influenced by aging in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the United States of America. 2002;99:10417–10422. doi: 10.1073/pnas.152161099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Morley JF, Morimoto RI. Regulation of longevity in Caenorhabditis elegans by heat shock factor and molecular chaperones. Molecular Biology of the Cell. 2004;15:657–664. doi: 10.1091/mbc.E03-07-0532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Morris JZ, Tissenbaum HA, Ruvkun G. A phosphatidylinositol-3-OH kinase family member regulating longevity and diapause in Caenorhabditis elegans. Nature. 1996;382:536–539. doi: 10.1038/382536a0. [DOI] [PubMed] [Google Scholar]
  99. Mostoslavsky R, Chua KF, Lombard DB, Pang WW, Fischer MR, Gellon L, Liu P, Mostoslavsky G, Franco S, Murphy MM, et al. Genomic instability and aging-like phenotype in the absence of mammalian SIRT6. Cell. 2006;124:315–329. doi: 10.1016/j.cell.2005.11.044. [DOI] [PubMed] [Google Scholar]
  100. Narendra D, Tanaka A, Suen DF, Youle RJ. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. Journal of Cell Biology. 2008;183:795–803. doi: 10.1083/jcb.200809125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Niedernhofer LJ, Garinis GA, Raams A, Lalai AS, Robinson AR, Appeldoorn E, Odijk H, Oostendorp R, Ahmad A, van Leeuwen W, 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]
  102. Nilsson EC, Long YC, Martinsson S, Glund S, Garcia-Roves P, Svensson LT, Andersson L, Zierath JR, Mahlapuu M. Opposite transcriptional regulation in skeletal muscle of AMP-activated protein kinase gamma3 R225Q transgenic versus knock-out mice. J Biol Chem. 2006;281:7244–7252. doi: 10.1074/jbc.M510461200. [DOI] [PubMed] [Google Scholar]
  103. Nisoli E, Tonello C, Cardile A, Cozzi V, Bracale R, Tedesco L, Falcone S, Valerio A, Cantoni O, Clementi E, et al. Calorie restriction promotes mitochondrial biogenesis by inducing the expression of eNOS. Science. 2005;310:314–317. doi: 10.1126/science.1117728. [DOI] [PubMed] [Google Scholar]
  104. Oberdoerffer P, Michan S, McVay M, Mostoslavsky R, Vann J, Park SK, Hartlerode A, Stegmuller J, Hafner A, Loerch P, et al. SIRT1 redistribution on chromatin promotes genomic stability but alters gene expression during aging. Cell. 2008;135:907–918. doi: 10.1016/j.cell.2008.10.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Olsen A, Vantipalli MC, Lithgow GJ. Lifespan extension of Caenorhabditis elegans following repeated mild hormetic heat treatments. Biogerontology. 2006;7:221–230. doi: 10.1007/s10522-006-9018-x. [DOI] [PubMed] [Google Scholar]
  106. Osler ME, Zierath JR. Adenosine 5'-monophosphate-activated protein kinase regulation of fatty acid oxidation in skeletal muscle. Endocrinology. 2008;149:935–941. doi: 10.1210/en.2007-1441. [DOI] [PubMed] [Google Scholar]
  107. Pan KZ, Palter JE, Rogers AN, Olsen A, Chen D, Lithgow GJ, Kapahi P. Inhibition of mRNA translation extends lifespan in Caenorhabditis elegans. Aging Cell. 2007;6:111–119. doi: 10.1111/j.1474-9726.2006.00266.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Pawlikowska L, Hu D, Huntsman S, Sung A, Chu C, Chen J, Joyner AH, Schork NJ, Hsueh WC, Reiner AP, et al. Association of common genetic variation in the insulin/IGF1 signaling pathway with human longevity. Aging cell. 2009;8:460–472. doi: 10.1111/j.1474-9726.2009.00493.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Pickford F, Masliah E, Britschgi M, Lucin K, Narasimhan R, Jaeger PA, Small S, Spencer B, Rockenstein E, Levine B, Wyss-Coray T. The autophagy-related protein beclin 1 shows reduced expression in early Alzheimer disease and regulates amyloid beta accumulation in mice. Journal of Clinical Investigation. 2008;118:2190–2199. doi: 10.1172/JCI33585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Powell CD, Quain DE, Smart KA. The impact of media composition and petite mutation on the longevity of a polyploid brewing yeast strain. Letters in applied microbiology. 2000;31:46–51. doi: 10.1046/j.1472-765x.2000.00766.x. [DOI] [PubMed] [Google Scholar]
  111. Qiang W, Weiqiang K, Qing Z, Pengju Z, Yi L. Aging impairs insulin-stimulated glucose uptake in rat skeletal muscle via suppressing AMPKalpha. Exp Mol Med. 2007;39:535–543. doi: 10.1038/emm.2007.59. [DOI] [PubMed] [Google Scholar]
  112. Ramsey KM, Mills KF, Satoh A, Imai S. Age-associated loss of Sirt1-mediated enhancement of glucose-stimulated insulin secretion in beta cell-specific Sirt1-overexpressing (BESTO) mice. Aging Cell. 2008;7:78–88. doi: 10.1111/j.1474-9726.2007.00355.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Rea SL, Ventura N, Johnson TE. Relationship between mitochondrial electron transport chain dysfunction, development, and life extension in Caenorhabditis elegans. PLoS Biol. 2007;5:e259. doi: 10.1371/journal.pbio.0050259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Richter K, Haslbeck M, Buchner J. Life on the verge of death: The heat shock response revisited. Mol. Cell. 2010;40:XXX. doi: 10.1016/j.molcel.2010.10.006. [DOI] [PubMed] [Google Scholar]
  115. Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, Puigserver P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature. 2005;434:113–118. doi: 10.1038/nature03354. [DOI] [PubMed] [Google Scholar]
  116. Rodier F, Coppe JP, Patil CK, Hoeijmakers WA, Munoz DP, Raza SR, Freund A, Campeau E, Davalos AR, Campisi J. Persistent DNA damage signalling triggers senescence-associated inflammatory cytokine secretion. Nat Cell Biol. 2009;11:973–979. doi: 10.1038/ncb1909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Rogina B, Helfand SL. Sir2 mediates longevity in the fly through a pathway related to calorie restriction. Proc Natl Acad Sci U S A. 2004;101:15998–16003. doi: 10.1073/pnas.0404184101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Ron D, Walter P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat Rev Mol Cell Biol. 2007;8:519–529. doi: 10.1038/nrm2199. [DOI] [PubMed] [Google Scholar]
  119. Rossi ML, Ghosh AK, Bohr VA. Roles of Werner syndrome protein in protection of genome integrity. DNA Repair (Amst) 2010;9:331–344. doi: 10.1016/j.dnarep.2009.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Saggerson D. Malonyl-CoA, a key signaling molecule in mammalian cells. Annual review of nutrition. 2008;28:253–272. doi: 10.1146/annurev.nutr.28.061807.155434. [DOI] [PubMed] [Google Scholar]
  121. Sasaki T, Maier B, Bartke A, Scrable H. Progressive loss of SIRT1 with cell cycle withdrawal. Aging Cell. 2006;5:413–422. doi: 10.1111/j.1474-9726.2006.00235.x. [DOI] [PubMed] [Google Scholar]
  122. Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, Coskun PE, Ladiges W, Wolf N, Van Remmen H, et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science. 2005;308:1909–1911. doi: 10.1126/science.1106653. [DOI] [PubMed] [Google Scholar]
  123. Schulz TJ, Zarse K, Voigt A, Urban N, Birringer M, Ristow M. Glucose restriction extends Caenorhabditis elegans life span by inducing mitochondrial respiration and increasing oxidative stress. Cell Metab. 2007;6:280–293. doi: 10.1016/j.cmet.2007.08.011. [DOI] [PubMed] [Google Scholar]
  124. Schwer B, Verdin E. Conserved metabolic regulatory functions of sirtuins. Cell Metab. 2008;7:104–112. doi: 10.1016/j.cmet.2007.11.006. [DOI] [PubMed] [Google Scholar]
  125. Sengupta S, Peterson TR, Sabatini DM. Regulation of the mTor complex 1 by nutrients, growth factors and stress. Mol Cell. 2010;40:XXXX. doi: 10.1016/j.molcel.2010.09.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Sharp ZD, Bartke A. Evidence for down-regulation of phosphoinositide 3-kinase/Akt/mammalian target of rapamycin (PI3K/Akt/mTOR)-dependent translation regulatory signaling pathways in Ames dwarf mice. J Gerontol A Biol Sci Med Sci. 2005;60:293–300. doi: 10.1093/gerona/60.3.293. [DOI] [PubMed] [Google Scholar]
  127. Shibata M, Lu T, Furuya T, Degterev A, Mizushima N, Yoshimori T, MacDonald M, Yankner B, Yuan J. Regulation of intracellular accumulation of mutant Huntingtin by Beclin 1. J Biol Chem. 2006;281:14474–14485. doi: 10.1074/jbc.M600364200. [DOI] [PubMed] [Google Scholar]
  128. Simonsen A, Cumming RC, Brech A, Isakson P, Schubert DR, Finley KD. Promoting basal levels of autophagy in the nervous system enhances longevity and oxidant resistance in adult Drosophila. Autophagy. 2008;4:176–184. doi: 10.4161/auto.5269. [DOI] [PubMed] [Google Scholar]
  129. Singh R, Kaushik S, Wang Y, Xiang Y, Novak I, Komatsu M, Tanaka K, Cuervo AM, Czaja MJ. Autophagy regulates lipid metabolism. Nature. 2009a;458:1131–1135. doi: 10.1038/nature07976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Singh R, Xiang Y, Wang Y, Baikati K, Cuervo AM, Luu YK, Tang Y, Pessin JE, Schwartz GJ, Czaja MJ. Autophagy regulates adipose mass and differentiation in mice. J Clin Invest. 2009b;119:3329–3339. doi: 10.1172/JCI39228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Sipula IJ, Brown NF, Perdomo G. Rapamycin-mediated inhibition of mammalian target of rapamycin in skeletal muscle cells reduces glucose utilization and increases fatty acid oxidation. Metabolism. 2006;55:1637–1644. doi: 10.1016/j.metabol.2006.08.002. [DOI] [PubMed] [Google Scholar]
  132. Sohal RS, Weindruch R. Oxidative stress, caloric restriction, and aging. Science. 1996;273:59–63. doi: 10.1126/science.273.5271.59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Spilman P, Podlutskaya N, Hart MJ, Debnath J, Gorostiza O, Bredesen D, Richardson A, Strong R, Galvan V. Inhibition of mTOR by rapamycin abolishes cognitive deficits and reduces amyloid-beta levels in a mouse model of Alzheimer's disease. PloS one. 2010;5:e9979. doi: 10.1371/journal.pone.0009979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Steffen KK, MacKay VL, Kerr EO, Tsuchiya M, Hu D, Fox LA, Dang N, Johnston ED, Oakes JA, Tchao BN, et al. Yeast life span extension by depletion of 60s ribosomal subunits is mediated by Gcn4. Cell. 2008;133:292–302. doi: 10.1016/j.cell.2008.02.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Suh Y, Atzmon G, Cho MO, Hwang D, Liu B, Leahy DJ, Barzilai N, Cohen P. Functionally significant insulin-like growth factor I receptor mutations in centenarians. Proc Natl Acad Sci U S A. 2008;105:3438–3442. doi: 10.1073/pnas.0705467105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Sun J, Tower J. FLP recombinase-mediated induction of Cu/Zn-superoxide dismutase transgene expression can extend the life span of adult Drosophila melanogaster flies. Mol Cell Biol. 1999;19:216–228. doi: 10.1128/mcb.19.1.216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Tissenbaum HA, Guarente L. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature. 2001;410:227–230. doi: 10.1038/35065638. [DOI] [PubMed] [Google Scholar]
  138. Trifunovic A, Wredenberg A, Falkenberg M, Spelbrink JN, Rovio AT, Bruder CE, Bohlooly YM, Gidlof S, Oldfors A, Wibom R, et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature. 2004;429:417–423. doi: 10.1038/nature02517. [DOI] [PubMed] [Google Scholar]
  139. Tyner SD, Venkatachalam S, Choi J, Jones S, Ghebranious N, Igelmann H, Lu X, Soron G, Cooper B, Brayton C, et al. p53 mutant mice that display early ageing-associated phenotypes. Nature. 2002;415:45–53. doi: 10.1038/415045a. [DOI] [PubMed] [Google Scholar]
  140. Um SH, Frigerio F, Watanabe M, Picard F, Joaquin M, Sticker M, Fumagalli S, Allegrini PR, Kozma SC, Auwerx J, Thomas G. Absence of S6K1 protects against age- and diet-induced obesity while enhancing insulin sensitivity. Nature. 2004;431:200–205. doi: 10.1038/nature02866. [DOI] [PubMed] [Google Scholar]
  141. Van Raamsdonk JM, Hekimi S. Deletion of the mitochondrial superoxide dismutase sod-2 extends lifespan in Caenorhabditis elegans. PLoS genetics. 2009;5:e1000361. doi: 10.1371/journal.pgen.1000361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Vermulst M, Bielas JH, Kujoth GC, Ladiges WC, Rabinovitch PS, Prolla TA, Loeb LA. Mitochondrial point mutations do not limit the natural lifespan of mice. Nature genetics. 2007;39:540–543. doi: 10.1038/ng1988. [DOI] [PubMed] [Google Scholar]
  143. Vermulst M, Wanagat J, Kujoth GC, Bielas JH, Rabinovitch PS, Prolla TA, Loeb LA. DNA deletions and clonal mutations drive premature aging in mitochondrial mutator mice. Nature genetics. 2008;40:392–394. doi: 10.1038/ng.95. [DOI] [PubMed] [Google Scholar]
  144. Weindruch R, Walford RL. The Retardation of Aging and Disease by Dietary Restriction. Springfield, Illinois: Charles C Thomas; 1988. [Google Scholar]
  145. Wellen KE, Thompson CB. Cellular metabolic stress: Considering what happens when cells overeat. Mol Cell. 2010;40:XXX. doi: 10.1016/j.molcel.2010.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Wenz T, Diaz F, Spiegelman BM, Moraes CT. Activation of the PPAR/PGC-1alpha pathway prevents a bioenergetic deficit and effectively improves a mitochondrial myopathy phenotype. Cell Metab. 2008;8:249–256. doi: 10.1016/j.cmet.2008.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  147. Wenz T, Rossi SG, Rotundo RL, Spiegelman BM, Moraes CT. Increased muscle PGC-1alpha expression protects from sarcopenia and metabolic disease during aging. Proc Natl Acad Sci U S A. 2009;106:20405–20410. doi: 10.1073/pnas.0911570106. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  148. Westerheide SD, Anckar J, Stevens SM, Jr, Sistonen L, Morimoto RI. Stress-inducible regulation of heat shock factor 1 by the deacetylase SIRT1. Science. 2009;323:1063–1066. doi: 10.1126/science.1165946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Willcox BJ, Donlan TA, He Q, Chen R, Grove JS, Yano K, Masaki KH, Willcox DC, Rodriguez B, Curb JD. FOXO3A genotype is strongly associated with human longevity. Proc Natl Acad Sci U S A. 2008;105:13987–13992. doi: 10.1073/pnas.0801030105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Wong E, Cuervo AM. Autophagy gone awry in neurodegenerative diseases. Nature Neuroscience. 2010;13:805–811. doi: 10.1038/nn.2575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Wullschleger S, Loewith R, Hall MN. TOR signaling in growth and metabolism. Cell. 2006;124:471–484. doi: 10.1016/j.cell.2006.01.016. [DOI] [PubMed] [Google Scholar]
  152. Yankner BA, Lu T, Loerch P. The Aging Brain. Annu Rev Pathol. 2008;3:41–66. doi: 10.1146/annurev.pathmechdis.2.010506.092044. [DOI] [PubMed] [Google Scholar]
  153. Yu WH, Cuervo AM, Kumar A, Peterhoff CM, Schmidt SD, Lee JH, Mohan PS, Mercken M, Farmery MR, Tjernberg LO, et al. Macroautophagy--a novel Beta-amyloid peptide-generating pathway activated in Alzheimer's disease. Journal of Cell Biology. 2005;171:87–98. doi: 10.1083/jcb.200505082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Yuan Z, Zhang X, Sengupta N, Lane WS, Seto E. SIRT1 regulates the function of the Nijmegen breakage syndrome protein. Molecular Cell. 2007;27:149–162. doi: 10.1016/j.molcel.2007.05.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Zhang C, Cuervo AM. Restoration of chaperone-mediated autophagy in aging liver improves cellular maintenance and hepatic function. Nat Med. 2008;14:959–965. doi: 10.1038/nm.1851. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES