J Neural Transm DOI 10.1007/s00702-013-1128-4

PSYCHIATRY AND PRECLINICAL PSYCHIATRIC STUDIES - REVIEW ARTICLE

Novel putative mechanisms to link circadian clocks to healthy aging Aurel Popa-Wagner • Bogdan Catalin Ana-Maria Buga



Received: 20 October 2013 / Accepted: 21 November 2013 Ó Springer-Verlag Wien 2013

Abstract The circadian clock coordinates the internal physiology to increase the homeostatic capacity thereby providing both a survival advantage to the system and an optimization of energy budgeting. Multiple-oscillator circadian mechanisms are likely to play a role in regulating human health and may contribute to the aging process. Our aim is to give an overview of how the central clock in the hypothalamus and peripheral clocks relate to aging and metabolic disorders, including hyperlipidemia and hyperglycemia. In particular, we unravel novel putative mechanisms to link circadian clocks to healthy aging. This review may lead to the design of large-scale interventions to help people stay healthy as they age by adjusting daily activities, such as feeding behavior, and or adaptation to agerelated changes in individual circadian rhythms. Keywords Clock genes  Aging  Tissue repair  Metabolic syndrome

Aging, a systems biology approach Aging is associated with a decline of locomotor, sensory, and cognitive performance in humans and experimental animals. This decline is partly due to the lifelong accumulation of damage at the molecular, cellular, tissue, and organ levels.

However, these insults alone cannot account for the systematic deterioration and loss of function that characterizes the senescent phenotype making it imperative for a multisystems, multidisciplinary, integrational approach to functional senescence. There are numerous theories of aging; however, no theory sufficiently explains all the changes of the aging process. Although some changes typically occur with aging, they occur at different rates and to different extents, i.e., we can speak of ‘‘mosaic aging’’ (Walker and Herndon 2010). To improve our understanding of cells and organisms as physiological, biochemical, and genetic systems, we have to study them as an integrated informational and metabolic system. Organismal senescence is the aging of whole organisms. Various types of damage have been proposed to accumulate with age, either due to an increased rate of production, or because of decreased repair or clearance of damage with time (Kirkwood 2005). However, none of these faults seem sufficiently numerous to cause the systematic deterioration and loss of function that characterizes the senescent phenotype (Kirkwood 2008). A first attempt to develop a ‘‘network theory of aging’’ that integrates the contributions of defective mitochondria, aberrant proteins, and free radicals in the aging process has been provided by Kirkwood and Kowald (1997).

Aging, autophagy, and clock genes A. Popa-Wagner (&)  A.-M. Buga Department of Psychiatry and Psychotherapy, University Rostock Medical School, Rostock, Germany e-mail: [email protected] A. Popa-Wagner  B. Catalin  A.-M. Buga Center of Clinical and Experimental Research, University of Medicine and Pharmacy, Craiova, Romania

Aging at the molecular level is characterized by increased levels of damaged proteins due to various post-translational modifications that include oxidation by free radicals of amino acid side chains, racemization of aspartyl and asparaginyl residues, deamidation of asparaginyl and glutaminyl residues, and oxidation of sulfhydryl groups.

123

A. Popa-Wagner et al.

There is extensive evidence that the increased concentration of oxidative damaged proteins can be a major contributing factor in aging (Ryazanov and Nefsky 2002). Protein misfolding is also likely to contribute to the increase in the concentration of abnormal enzymes and proteins in senescent tissues. To avoid aging, the cell structure within tissues must be maintained by a continuous synthesis and degradation of worn-out molecules process, i.e., a continuous turnover process. Efficient macromolecular turnover is integral to the normal function and survival of a biological system. Although there are large variations in the rates of degradation of individual proteins, it is generally observed that the overall protein turnover slows down during aging. The physiological consequences of decreased protein turnover include the accumulation of altered and abnormal proteins, an altered pattern of posttranslational modifications due to increased dwell time and a disruption of the organization of the cytoskeleton and extracellular matrix (Martinez-Vicente et al. 2005; Rattan 2010). One mechanism involved in the regeneration of metabolic precursors and clearing subcellular debris is autophagy. Autophagy participates in the turnover of organelle (mitochondria, endoplasmic reticulum, and peroxisomes) and is required for the clearance of oxidized proteins and removal of insoluble protein aggregates which accumulate during aging and disease. Indeed, modulators of autophagy-like caloric restriction or rapamycin lowers mTOR signaling and extends the average and maximum life span of a variety of organisms including yeast, flies, worms, fish, and rodents (Fabrizio et al. 2001; Fontana et al. 2010). Furthermore, the level of autophagy genes, such as ATG5, ATG7, and BECN1 decline in the aged brain (Choi et al. 2013). Therefore, the homeostatic functions of autophagy with respect to turnover of long-lived proteins and removal of damaged organelles, and cellular debris are believed to constitute an antiaging process (Rubinsztein et al. 2011). All tissular activities are under central nervous system (CNS) control. A recent study has shown that in the lung, more than two-thirds of the probe sets showing cyclic expression peaked during the animal’s light/inactive period. Many of the genes that peaked during the inactive period included genes related to extracellular matrix, cytoskeleton, and macromolecular trafficking, which appear to be mainly involved in the turnover and remodelling of the organ (Sukumaran et al. 2011). At the same time, light triggers coordinated cyclic AMP-response element-binding protein expression and mTOR activation in the suprachiasmatic nucleus (SCN) and neurons (Cao et al. 2008). Conversely, intraventricular infusion of the mTOR inhibitor rapamycin, led to a significant attenuation of the phase-delaying effect of early-night light, while abrogation of mTOR signaling led to a significant attenuation of light-

123

evoked PERIOD protein expression (Cao and Obrietan 2010).

Clock genes and aging European Union societies are all facing the effects of aging populations and the huge costs associated with age-related illnesses. Therefore, promoting healthy aging by adjusting daily activities, such as feeding behavior or adaptation to age-related changes in individual circadian rhythms is fundamental for the sustainable development of our society (Levi and Schibler 2007). The discovery of the molecular clocks that control cell activity has expanded our understanding of physiology and behavior. However, the rhythmic mechanisms by which transcriptional signals can coordinate behavioral and physiological processes are not fully elucidated (Cho et al. 2012). Anatomically, the circadian rhythms are organized in a hierarchical cascade starting from the CNS and ending in the periphery. Centrally, the circadian clock is encoded by master pacemaker neurons localized in the SCN. The input signal to the master pacemaker neurons of the SCN is provided by the light/dark cycle via the retina (Hastings et al. 2003). The circadian rhythm is driven by a molecular clock involving a transcriptional negative-feedback loop. Upon day-light stimulation, the E-box containing transcriptional activators BMAL1/NPAS2 and CLOCK will stimulate the activity of Period and Cryptochrome genes causing in turn an increase in the abundance of PER1-3 and CRY1-2 repressors that will then accumulate during the day. The PER1-3 and CRY1-2 repressors will have a negative feedback on CLOCK:NPAS2/BMAL1 transcription complex. This mechanism creates an autoregulatory loop that controls all circadian rhythms (Kume et al. 1999; Lee et al. 2001; Huang et al. 2011). The master clock in the SCN sends signals to the extraSCN neurons, which in turn modulate behavioral and physiological rhythms by aligning circadian gene oscillations within both extra-SCN neurons and peripheral tissue (Weaver 1998; Huang et al. 2011). In the peripheral tissues, local physiology is rhythmically regulated by tissue-specific networks of ligand-activated transcriptional factors that display with little or no overlap across tissues in humans (Fig. 1) and rodents (Fig. 2) (Yang et al. 2006; Hughes et al. 2009). Evolutionarily, the synchronization of circadian clocks provides an adaptive advantage by enhancing an organism’s ability to respond to daily changes in light, temperature and humidity (circadian resonance) (Woelfle et al. 2004). In invertebrates, the circadian clock coordinates the internal physiology to increase the homeostatic capacity,

Circadian clocks to healthy aging Fig. 1 Circadian rhythmicity of Bmal 1, Per3, and DBP in the human CNS. Circadian rhythmicity for physiological markers of body clock (cortisol, melatonin, temperature, and blood pressure) is also shown

thereby providing both a survival advantage to the system and an optimization of energy budgeting (Pittendrigh and Minis 1972; von Saint and Aschoff 1978; Dodd et al. 2005). In mammals, for example, a nocturnal rodent possessing a circadian timekeeper can anticipate dusk in his underground habitat and does not have to forage periodically to examine whether sunset is approaching and such anticipation may considerably reduce exposure to active daytime predators. However, with increasing age, the circadian rhythm becomes more and more disrupted. We hypothesize that age-associated disruptions in the central circadian oscillator will diminish the systemic advantage and accelerate the aging process. The hypothesized link between the circadian clock, aging, and systemic advantage is supported by several lines of evidence. A recent study investigating the association between mammalian lifespan and endogenous circadian free-running period (tau), reported that deviation of tau from 24 h was inversely related to the lifespans of rodent and primate species (Wyse et al. 2010). This result suggests that organismal efforts to redress the misalignment of endogenous rhythms and 24 h environmental cycles may be associated with a physiological cost that has a negative impact on longevity (Wyse et al. 2010). Indeed, lack of re-entrainment of endogenous circadian clock due to defective melatonin synthesis or melatonin receptors in naked mole rat leads to extended life span (Kim et al. 2011). Aging may result in selective reduced expression of the genes encoding the core clock genes like Bmal1, in the SCN, cortex, hippocampus, and caudate putamen or altered Per2 oscillations in the CA1, DG, dorsomedial

hypothalamus, and piriform cortex (Duncan et al. 2013). Moreover, Bmal1-/- mice have reduced life span and develop various symptoms of premature aging like tissue atrophy and cognitive deficits that have been largely attributed to increased tissular level of reactive oxygen species (Kondratov et al. 2006). Age-related changes in circadian function are accelerated in Alzheimer disease (AD) and are most likely due to neurodegeneration of SCN cells (Cermakian et al. 2011; Bonaconsa et al. 2013; Coogan et al. 2013). AD patients exhibit less diurnal and increased nocturnal activity that becomes manifested by nocturnal awakening and increased daytime sleep bouts (McClung 2011). Similarly, the rhythmic expression of Bmal1, Cry1, Per1 in the pineal gland of AD is lost (Wu et al. 2007). Similar alterations in sleep–wake rhythms, but no significant changes in the daily activity rhythms have been recorded in the APP 9 PS1 mouse model of AD (Duncan et al. 2012). Along the same line of evidence, injection of aggregates of Abeta or cells overexpressing the toxic Abeta into the SCN of rodents causes alterations in circadian phase and a decrease in amplitude (Tate et al. 1992; Furio et al. 2002).

Circadian rhythm and health Disturbances in the circadian rhythm may have dramatic effects on our health. Circadian disruption is associated with increased morbidity and mortality (Kubo et al. 2006; Gery and Koeffler 2007), whereas robust and reset circadian rhythms could lead to better health and increased longevity (Gibson et al. 2009; Froy 2011). Moreover, phase

123

A. Popa-Wagner et al. Fig. 2 Circadian rhythmicity of Clock and Bmal in the CNS and various peripheral rodent organs along with circadian variation of organ-specific genes, such as Scd1 and Srebf1 (liver), Cpl1b and Falp1 (striated muscle), Dgat1 and Mgat (fat tissue), Glut4 and Hlf (heart), Dbp (smooth muscle), Pai1 (blood vessels), and Tim (pancreas). Circadian rhythmicity for physiological markers of body clock (cortisol, melatonin, temperature, and blood pressure) is also shown. Please note the mirrored expression pattern for Bmal in the striated muscle, heart, blood vessels, and pancreas vs SNC suggesting that in the periphery each organ has a specific expression pattern

123

Circadian clocks to healthy aging

shifts of the light/dark cycle induced significant mortality in aged animals (Davidson et al. 2006). It is well documented that the circadian rhythm becomes disorganized with increasing age (Kolker et al. 2003). Therefore, it comes at no surprise that transgenic mice exhibiting regular circadian patterns of food intake and body temperature during aging have increased life spans (Froy et al. 2006).

Circadian rhythm and metabolic disorders Shift workers are at an unusually high risk of metabolic syndrome, obesity, and diabetes (Bass and Takahashi 2010; Spiegel et al. 2009), not because of the nature of shift work, but because of the resulting disruptions of the circadian rhythms of meals and sleep (Pulivarthy et al. 2007). Rhythmic oscillations with a 24 h period for insulin, glucose, and leptin levels are well known (Sinha et al. 1996; Dupuis et al. 2010). Furthermore genome-wide association studies have linked clock gene variation to fasting glucose levels (Woon et al. 2007; Dupuis et al. 2010), obesity and metabolic syndrome (Scott et al. 2008). Many other studies have showed that aging-associated diseases, such as diabetes, metabolic syndrome, sleep disorders (Van Cauter et al. 1997; Bass and Turek 2005), narcolepsy (Howell et al. 2009), and obesity (Taheri et al. 2004; Spiegel et al. 1999) are tightly linked to the disturbances of the environmental light/dark cycle circadian rhythm. Mice and humans with disrupted circadian rhythms have been shown to develop metabolic disorders, including hyperlipidemia and hyperglycemia, suggesting a link between clock function and metabolic functions (Huang et al. 2011). The link is provided by an accessory feedback loop made of two nuclear receptors REV-ERB-a and REVERB-b that contribute primarily to clock stabilization. In addition, REV-ERB-a and REV-ERB-b have been also proposed to have a role in the regulation of energy metabolism (Levi and Schibler 2007; Preitner et al. 2002; Liu et al. 2008). The nuclear receptor REV-ERB alpha plays a key role in lipid metabolism and adipocyte differentiation, hepatic gluconeogenesis, and Bmal1 transcription (Yin et al. 2007; Han et al. 2012). Moreover, double knockout Rev-erb-a and Rev-erb-b mice displayed markedly altered circadian wheel running behavior and deregulated lipid homeostatic gene networks (Cho et al. 2012). In peripheral tissue like the liver, circadian oscillators may help to respond to feeding-fasting cycles by separating glycogen synthesis and degradation in time, limiting the former to the absorptive phase and the latter to the postabsorptive phase (Luna-Moreno et al. 2012). It has been demonstrated, at molecular level, that interactions between

clock networks and peripheral nutrient sensors are bidirectional: acute nutrient withdrawal or caloric restriction will induce NAD?-dependent deacetylase sirtuin 1 (SIRT1), a gene heavily involved in the regulation of life span in lower organisms and aging (Nemoto et al. 2004; Cohen et al. 2004). SIRT1 in turn can bind to CLOCK/ BMAL1 complex to regulate expression of clock genes (Asher et al. 2008; Nakahata et al. 2009; Ramsey et al. 2009). Conversly, the Clock gene can also directly regulate the energetic status of a cell by modulating the NAD? circadian oscillations and closing the regulatory loop (Nakahata et al. 2009; Ramsey et al. 2009). Recently, it was shown that Sirt1 regulates central circadian control in the brains of mice to determine the period, activity levels, and ability to adjust to re-entrainment of the central circadian clock by amplying the activity of Bmal1 (Chang and Guarente 2013). SIRT1 is an NADdependent protein deacetylase that is involved in aging and energy metabolism (Finkel et al. 2009). SIRT1 mediates the beneficial effects of caloric restriction, and its activity is thus critical in the maintenance of health (Guarente 2012; Chang and Guarente 2013).

Modulators of clock genes expression Orphan nuclear receptor ROR alpha and Rev-erba are output regulators of the circadian clock that target tissue metabolic functions (Lin et al. 2008). The loop is closed by peroxisome proliferator-activated receptors (PPAR) which are sensors of intracellular lipids and regulate lipid synthesis, storage, and fatty acid oxidation, and convey signals to the central clock by interfering with the expression of Bmal1/Clock. The nuclear receptors PPARa and c bind to the Reverba and Bmal1 promoters and up-regulate their expression. Finally, PPARc co-activator (PGC)1a potentiates RORa transcriptional activity and enhances Rev-erba and Bmal1 transcription (Lin et al. 2005). As a result, the PPAR alpha and PPAR gamma can compete with REV-ERB alpha in binding to Bmal1 promoter and inducing Bmal1 expression (Sato et al. 2004). PPAPR transcriptional factor has different roles depending on its subtype. PPAR gamma induces Bmal1 expression in blood vessels and has an important role in adipocyte differentiation and triglyceride synthesis (Semple et al. 2006). PPARa stimulates fatty acids oxidation and regulates genes controlling lipid homeostasis and as such may prevent atherosclerosis (Canaple et al. 2006). It has also been showed that SIRT1, a sensor of the energetic status of a cell, can influence clock networks through multiple mechanisms by suppressing PPAR gamma activity (Picard et al. 2004) or activating Peroxisome

123

A. Popa-Wagner et al.

proliferator-activated receptor-gamma coactivator (Rodgers et al. 2005; Huang et al. 2011). Such studies on the circadian clock have opened new ways to treat metabolic and brain diseases with huge impact in the fields of cardiology, neurology and psychiatry (Solt et al. 2011). For example, REV-ERB alpha expression has been showed to be affected via glycogen synthase kinase 3 by lithium which is used to treat bipolar disorder, thus establishing Rev-erb alpha as a critical component of the peripheral clock and a biological target of lithium therapy (Yin et al. 2006). Other animal studies have showed that synthetic REV-ERB agonists can decrease fat mass in mice that where diet-induced to become obese, and also reduced dyslipidemia and hyperglycemia syndromes (Solt et al. 2012).

Conclusion Further research is required to understand how the master clock responds to environmental changes, such as food intake, light/dark and sleep/awake cycles, and how these affect peripheral clocks. In the end we shall be able to develop target-specific drugs that specifically interfere with the master clock or peripheral tissue clocks. Acknowledgments This work was supported by UEFISCDI, PN-IIID-PCE-2011-3-0848 and UEFISCDI FLARE2 (to A.M.B.). Conflict of interest

None.

References Asher G, Gatfield D, Stratmann M et al (2008) SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell 134:317–328 Bass J, Takahashi JS (2010) Circadian integration of metabolism and energetics. Science 330:1349–1354 Bass J, Turek FW (2005) Sleepless in America: a pathway to obesity and the metabolic syndrome? Arch Intern Med 165:15–16 Bonaconsa M, Colavito V, Pifferi F, Aujard F, Schenker E, Dix S, Grassi-Zucconi G, Bentivoglio M, Bertini G (2013) Cell clocks and neuronal networks: neuron ticking and synchronization in aging and aging-related neurodegenerative disease. Curr Alzheimer Res 10:597–608 Canaple L, Rambaud J, Dkhissi-Benyahya O et al (2006) Reciprocal regulation of brain and muscle Arnt-like protein 1 and peroxisome proliferator-activated receptor alpha defines a novel positive feedback loop in the rodent liver circadian clock. Mol Endocrinol 20:1715–1727 Cao R, Obrietan K (2010) mTOR signaling and entrainment of the mammalian circadian Clock. Mol Cell Pharmacol 2:125–130 Cao R, Lee B, Cho HY, Saklayen S, Obrietan K (2008) Photic regulation of the mTOR signaling pathway in the suprachiasmatic circadian clock. Mol Cell Neurosci 38:312–324

123

Cermakian N, Lamont EW, Boudreau P, Boivin DB (2011) Circadian clock gene expression in brain regions of Alzheimer’s disease patients and control subjects. J Biol Rhythm 26:160–170 Chang HC, Guarente L (2013) SIRT1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell 153:1448–1460 Cho H, Zhao X, Hatori M et al (2012) Regulation of circadian behaviour and metabolism by REV-ERB-a and REV-ERB-b. Nature 485:123–127 Choi AM, Ryter SW, Levine B (2013) Autophagy in human health and disease. N Engl J Med 368:651–662 Cohen HY, Miller C, Bitterman KJ et al (2004) Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase. Science 305:390–392 Coogan AN, Schutova´ B, Husung S, Furczyk K, Baune BT, Kropp P, Ha¨ßler F, Thome J (2013) The circadian system in Alzheimer’s disease: disturbances, mechanisms, and opportunities. Biol Psychiatry 74:333–339 Davidson AJ, Sellix MT, Daniel J, Yamazaki S, Menaker M, Block GD (2006) Chronic jet-lag increases mortality in aged mice. Curr Biol 16:R914–R916 Dodd AN, Salathia N, Hall A, Ke´vei E, To´th R, Nagy F, Hibberd JM, Millar AJ, Webb AA (2005) Plant circadian clocks increase photosynthesis, growth, survival, and competitive advantage. Science 309(5734):630–633 Duncan MJ, Smith JT, Franklin KM, Beckett TL, Murphy MP, St Clair DK, Donohue KD, Striz M, O’Hara BF (2012) Effects of aging and genotype on circadian rhythms, sleep, and clock gene expression in APP 9 PS1 knock-in mice, a model for Alzheimer’s disease. Exp Neurol 236:249–258 Duncan MJ, Prochot JR, Cook DH, Tyler Smith J, Franklin KM (2013) Influence of aging on Bmal1 and Per2 expression in extra-SCN oscillators in hamster brain. Brain Res 1491:44–53 Dupuis J, Langenberg C, Prokopenko I et al (2010) New genetic loci implicated in fasting glucose homeostasis and their impact on type 2 diabetes risk. Nat Genet 42:105–116 Fabrizio P, Pozza F, Pletcher SD, Gendron CM, Longo VD (2001) Regulation of longevity and stress resistance by Sch9 in yeast. Science 292:288–290 Finkel T, Deng CX, Mostoslavsky R (2009) Recent progress in the biology and physiology of sirtuins. Nature 460:587–591 Fontana L, Partridge L, Longo VD (2010) Extending healthy life span—from yeast to humans. Science 328:321–326 Froy O (2011) Circadian rhythms, aging and life span in mammals. Physiol Bethesda 26:225–235 Froy O, Chapnik N, Miskin R (2006) Long-lived al-phaMUPA transgenic mice exhibit pronounced circadian rhythms. Am J Physiol Endocrinol Metab 291:E1017–E1024 Furio AM, Cutrera RA, Castillo Thea V et al (2002) Effect of melatonin on changes in locomotor activity rhythm of Syrian hamsters injected with beta amyloid peptide 25–35 in the suprachiasmatic nuclei. Cell Mol Neurobiol 22:699–709 Gery S, Koeffler HP (2007) The role of circadian regulation in cancer. Cold Spring Harb Symp Quant Biol 72:459–464 Gibson EM, Williams WP, Kriegsfeld LJ (2009) Aging in the circadian system: considerations for health, disease prevention and longevity. Exp Gerontol 44:51–56 Guarente L (2012) Sirtuins and calorie restriction. Nat Rev Mol Cell Biol 13:207. doi:10.1038/nrm3308 Han C, Zhao X, Hatori M et al (2012) Evans regulation of circadian behaviour and metabolism by REV-ERB-a and REV-ERB-b. Nature 485:123–127 Hastings MH, Reddy AB, Maywood ES (2003) A clockwork web: circadian timing in brain and periphery, in health and disease. Nat Rev Neurosci 4:649–661

Circadian clocks to healthy aging Howell MJ, Schenck CH, Crow SJ (2009) A review of nighttime eating disorders. Sleep Med Rev 13:23–34 Huang W, Ramsey KM, Marcheva B, Bass J (2011) Circadian rhythms, sleep, and metabolism. J Clin Invest 121:2133–2141 Hughes ME, DiTacchio L, Hayes KR et al (2009) Harmonics of circadian gene transcription in mammals. PLoS Genet 5:e1000442 Kim EB, Fang X, Fushan AA et al (2011) Genome sequencing reveals insights into physiology and longevity of the naked mole rat. Nature 479:223–227 Kirkwood TB (2005) Understanding the odd science of aging. Cell 120:437–447 Kirkwood TB (2008) A systematic look at an old problem. Nature 451:644–647 Kirkwood TB, Kowald A (1997) Network theory of aging. Exp Gerontol 32:395–399 Kolker DE, Fukuyama H, Huang DS, Takahashi JS, Horton TH, Turek FW (2003) Aging alters circadian and light-induced expression of clock genes in golden hamsters. J Biol Rhythm 18:159–169 Kondratov RV, Kondratova AA, Gorbacheva VY, Vykhovanets OV, Antoch MP (2006) Early aging and age-related pathologies in mice deficient in BMAL1, the core component of the circadian clock. Genes Dev 20:1868–1873 Kubo T, Ozasa K, Mikami K (2006) Prospective cohort study of the risk of prostate cancer among rotating-shift workers: findings from the Japan collaborative cohort study. Am J Epidemiol 164:549–555 Kume K, Zylka MJ, Sriram S et al (1999) mCRY1 and mCRY2 are essential components of the negative limb of the circadian clock feedback loop. Cell 98:193–205 Lee C, Etchegaray JP, Cagampang FR, Loudon AS, Reppert SM (2001) Posttranslational mechanisms regulate the mammalian circadian clock. Cell 107:855–867 Levi F, Schibler U (2007) Circadian rhythms: mechanism and therapeutic implications. Annu Rev Pharmacol 47:493–528 Lin J, Handschin C, Spiegelman BM (2005) Metabolic control through the PGC-1 family of transcription coactivators. Cell Metab 1:361–370 Lin JD, Liu C, Li S (2008) Integration of energy metabolism and the mammalian clock. Cell Cycle 7:453–457 Liu AC, Tran HG, Zhang EE, Priest AA, Welsh DK, Kay SA (2008) Redundant function of REV-ERBalpha and beta and nonessential role for Bmal1 cycling in transcriptional regulation of intracellular circadian rhythms. PLoS Genet 4:e1000023. doi:10. 1371/journal.pgen.1000023 Luna-Moreno D, Garcı´a-Ayala B, Dı´az-Mun˜oz M (2012) Daytime restricted feeding modifies 24 h rhythmicity and subcellular distribution of liver glucocorticoid receptor and the urea cycle in rat liver. Br J Nutr 1:12 Martinez-Vicente M, Sovak G, Cuervo AM (2005) Protein degradation and aging. Exp Gerontol 40:622–633 McClung CA (2011) Circadian rhythms and mood regulation: insights from pre-clinical models. Eur Neuropsychopharmacol 21(Suppl 4):S683–S693 Nakahata Y, Sahar S, Astarita G, Kaluzova M, Sassone-Corsi P (2009) Circadian control of the NAD? salvage pathway by CLOCK-SIRT1. Science 324:654–657 Nemoto S, Fergusson MM, Finkel T (2004) Nutrient availability regulates SIRT1 through a forkhead-dependent pathway. Science 306:2105–2108 Picard F, Kurtev M, Chung N et al (2004) Sirt1 promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 429:771–776 Pittendrigh CS, Minis DH (1972) Circadian systems: longevity as a function of circadian resonance in Drosophila melanogaster. Proc Natl Acad Sci USA 69:1537–1539

Preitner N, Damiola F, Lopez-Molina L, Zakany J, Duboule D, Albrecht U, Schibler U (2002) The orphan nuclear receptor REV-ERBalpha controls circadian transcription within the positive limb of the mammalian circadian oscillator. Cell 110:251–260 Pulivarthy SR, Tanaka N, Welsh DK, De Haro L, Verma IM, Panda S (2007) Reciprocity between phase shifts and amplitude changes in the mammalian circadian clock. PNAS 104:20356–20361 Ramsey KM, Yoshino J, Brace CS et al (2009) Circadian clock feedback cycle through NAMPT-mediated NAD? biosynthesis. Science 324:651–654 Rattan SI (2010) Synthesis, modification and turnover of proteins during aging. Adv Exp Med Biol 694:1–13 Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, Puigserver P (2005) Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 434:113–118 Rubinsztein DC, Marin˜o G, Kroemer G (2011) Autophagy and aging. Cell 2011(146):682–695 Ryazanov AG, Nefsky BS (2002) Protein turnover plays a key role in aging. Mech Ageing Dev 123:207–213 Sato TK, Panda S, Miraglia LJ et al (2004) A functional genomics strategy reveals Rora as a component of the mammalian circadian clock. Neuron 43:527–537 Scott EM, Carter AM, Grant PJ (2008) Association between polymorphisms in the Clock gene, obesity and the metabolic syndrome in man. Int J Obes Lond 32:658–662 Semple RK, Chatterjee VK, O’Rahilly S (2006) PPAR gamma and human metabolic disease. J Clin Invest 116:581–589 Sinha MK, Ohannesian JP, Heiman ML et al (1996) Nocturnal rise of leptin in lean, obese, and non-insulin-dependent diabetes mellitus subjects. J Clin Invest 97:1344–1347 Solt LA, Kojetin DJ, Burris TP (2011) The REV-ERBs and RORs: molecular links between circadian rhythms and lipid homeostasis. Future Med Chem 3:623–638 Solt LA, Wang Y, Banerjee S et al (2012) Regulation of circadian behavior and metabolism by synthetic REV-ERB agonists. Nature 485:62–68 Spiegel K, Leproult R, Van Cauter E (1999) Impact of sleep debt on metabolic and endocrine function. Lancet 354:1435–1439 Spiegel K, Tasali E, Leproult R, Van Cauter E (2009) Effects of poor and short sleep on glucose metabolism and obesity risk. Nat Rev Endocrinol 5:253–261 Sukumaran S, Jusko WJ, DuBois DC, Almon RR (2011) Light–dark oscillations in the lung transcriptome: implications for lung homeostasis, repair, metabolism, disease, and drug action. J Appl Physiol 110:1732–1747 Taheri S, Lin L, Austin D, Young T, Mignot E (2004) Short sleep duration is associated with reduced leptin, elevated ghrelin, and increased body mass index. PLoS Med 1:e3 Tate B, Aboody-Guterman KS, Morris AM, Walcott EC, Majocha RE, Marotta CA (1992) Disruption of circadian regulation by brain grafts that overexpress Alzheimer beta/A4 amyloid. Proc Natl Acad Sci USA 89:7090–7094 Van Cauter E, Polonsky KS, Scheen AJ (1997) Roles of circadian rhythmicity and sleep in human glucose regulation. Endocr Rev 18:716–738 von Saint Paul U, Aschoff J (1978) Longevity among blowflies Phormia terraenovae R. D. kept in non-24-hour light–dark cycles. J Comp Physiol 127:191–195 Walker LC, Herndon JG (2010) Mosaic aging. Med Hypotheses 74:1048–1051 Weaver DR (1998) The suprachiasmatic nucleus: a 25-year retrospective. J Biol Rhythms 13:100–112 Woelfle MA, Ouyang Y, Phanvijhitsiri K, Johnson CH (2004) The adaptive value of circadian clocks: an experimental assessment in cyanobacteria. Curr Biol 14:1481–1486

123

A. Popa-Wagner et al. Woon PY, Kaisaki PJ, Braganc¸a J et al (2007) Aryl hydrocarbon receptor nuclear translocator-like (BMAL1) is associated with susceptibility to hypertension and type 2 diabetes. Proc Natl Acad Sci USA 104:14412–14417 Wu YH, Zhou JN, Van Heerikhuize J, Jockers R, Swaab DF (2007) Decreased MT1 melatonin receptor expression in the suprachiasmatic nucleus in aging and Alzheimer’s disease. Neurobiol Aging 28:1239–1247 Wyse CA, Coogan AN, Selman C, Hazlerigg DG, Speakman JR (2010) Association between mammalian lifespan and circadian

123

free-running period: the circadian resonance hypothesis revisited. Biol Lett 6:696–698 Yang X, Downes M, Yu RT et al (2006) Nuclear receptor expression links the circadian clock to metabolism. Cell 126:801–810 Yin L, Wang J, Klein PS, Lazar MA (2006) Nuclear receptor Rev-erb alpha is a critical lithium-sensitive component of the circadian clock. Science 311:1002–1005 Yin L, Wu N, Curtin JC et al (2007) Rev-erb alpha, a heme sensor that coordinates metabolic and circadian pathways. Science 318:1786–1789

Novel putative mechanisms to link circadian clocks to healthy aging.

The circadian clock coordinates the internal physiology to increase the homeostatic capacity thereby providing both a survival advantage to the system...
400KB Sizes 0 Downloads 0 Views