Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

The brain, sirtuins, and ageing

Key Points

  • Various cellular and molecular changes occur in the constituents of the brain (for example, neurons, microglia, astrocytes and oligodendrocytes) during the ageing process.

  • Functional alterations, such as cognitive decline, mental deficits, sleep disruption and circadian dysfunction, are induced by these cellular and molecular changes.

  • Sirtuins have crucial roles in the regulation of brain function during the ageing process.

  • Activation of microglia, possibly caused by reduced sirtuin activity, induces prolonged neuroinflammation, resulting in synaptic damage and neuronal death.

  • Age-associated decline in hypothalamic function, which is probably due to reduced SIRT1 activity, mediates ageing at a systemic level and ultimately affects longevity in mammals.

  • Peripheral tissues communicate with the brain via a range of hormones and circulating factors, which have been characterized by parabiosis experiments, thus affecting the pathophysiology of brain ageing.

Abstract

In mammals, recent studies have demonstrated that the brain, the hypothalamus in particular, is a key bidirectional integrator of humoral and neural information from peripheral tissues, thus influencing ageing both in the brain and at the 'systemic' level. CNS decline drives the progressive impairment of cognitive, social and physical abilities, and the mechanisms underlying CNS regulation of the ageing process, such as microglia–neuron networks and the activities of sirtuins, a class of NAD+-dependent deacylases, are beginning to be understood. Such mechanisms are potential targets for the prevention or treatment of age-associated dysfunction and for the extension of a healthy lifespan.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Cellular, histological and functional changes in normal brain ageing.
Figure 2: Exacerbation of brain ageing through the activation of microglia.
Figure 3: Communication between the brain and peripheral tissues.

Similar content being viewed by others

References

  1. Plassman, B. L. et al. Prevalence of cognitive impairment without dementia in the United States. Ann. Intern. Med. 148, 427–434 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Krell-Roesch, J. et al. FDG-PET and neuropsychiatric symptoms among cognitively normal elderly persons: the Mayo Clinic Study of Aging. J. Alzheimers Dis. 53, 1609–1616 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Piovezan, R. D. et al. The impact of sleep on age-related sarcopenia: possible connections and clinical implications. Ageing Res. Rev. 23, 210–220 (2015).

    Article  PubMed  Google Scholar 

  4. Collier, T. J., Kanaan, N. M. & Kordower, J. H. Ageing as a primary risk factor for Parkinson's disease: evidence from studies of non-human primates. Nat. Rev. Neurosci. 12, 359–366 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Guarente, L. Aging research — where do we stand and where are we going? Cell 159, 15–19 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Satoh, A. & Imai, S. Systemic regulation of mammalian ageing and longevity by brain sirtuins. Nat. Commun. 5, 4211 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Norden, D. M. & Godbout, J. P. Review: microglia of the aged brain: primed to be activated and resistant to regulation. Neuropathol. Appl. Neurobiol. 39, 19–34 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Lull, M. E. & Block, M. L. Microglial activation and chronic neurodegeneration. Neurotherapeutics 7, 354–365 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Imai, S. & Guarente, L. It takes two to tango: NAD+ and sirtuins in aging/longevity control. NPJ Aging Mech. Dis. 2, 16017 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Kanfi, Y. et al. The sirtuin SIRT6 regulates lifespan in male mice. Nature 483, 218–221 (2012).

    Article  CAS  PubMed  Google Scholar 

  11. Satoh, A. et al. Sirt1 extends life span and delays aging in mice through the regulation of Nk2 homeobox 1 in the DMH and LH. Cell Metab. 18, 416–430 (2013). This study demonstrates that the mammalian NAD+-dependent deacetylase SIRT1 plays a crucial part in the regulation of mammalian ageing and longevity through a hypothalamic pathway.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Igarashi, M. & Guarente, L. mTORC1 and SIRT1 cooperate to foster expansion of gut adult stem cells during calorie restriction. Cell 166, 436–450 (2016).

    Article  CAS  PubMed  Google Scholar 

  13. Russell, S. J. & Kahn, C. R. Endocrine regulation of ageing. Nat. Rev. Mol. Cell Biol. 8, 681–691 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Hong, S., Zhao, B., Lombard, D. B., Fingar, D. C. & Inoki, K. Cross-talk between sirtuin and mammalian target of rapamycin complex 1 (mTORC1) signaling in the regulation of S6 kinase 1 (S6K1) phosphorylation. J. Biol. Chem. 289, 13132–13141 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ramadori, G. et al. SIRT1 deacetylase in POMC neurons is required for homeostatic defenses against diet-induced obesity. Cell Metab. 12, 78–87 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Ramadori, G. et al. SIRT1 deacetylase in SF1 neurons protects against metabolic imbalance. Cell Metab. 14, 301–312 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wu, D., Qiu, Y., Gao, X., Yuan, X. B. & Zhai, Q. Overexpression of SIRT1 in mouse forebrain impairs lipid/glucose metabolism and motor function. PLoS ONE 6, e21759 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Peters, A. Structural changes that occur during normal aging of primate cerebral hemispheres. Neurosci. Biobehav. Rev. 26, 733–741 (2002).

    Article  PubMed  Google Scholar 

  19. Price, J. L. & Morris, J. C. Tangles and plaques in nondemented aging and “preclinical” Alzheimer's disease. Ann. Neurol. 45, 358–368 (1999).

    Article  CAS  PubMed  Google Scholar 

  20. Petralia, R. S., Mattson, M. P. & Yao, P. J. Communication breakdown: the impact of ageing on synapse structure. Ageing Res. Rev. 14, 31–42 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Seib, D. R. & Martin-Villalba, A. Neurogenesis in the normal ageing hippocampus: a mini-review. Gerontology 61, 327–335 (2015).

    Article  CAS  PubMed  Google Scholar 

  22. Alvarez-Buylla, A. & Garcia-Verdugo, J. M. Neurogenesis in adult subventricular zone. J. Neurosci. 22, 629–634 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Pignatelli, A. & Belluzzi, O. in The Neurobiology of Olfaction (ed. Menini, A.) (CRC Press, 2010).

    Google Scholar 

  24. Gould, E. et al. Hippocampal neurogenesis in adult Old World primates. Proc. Natl Acad. Sci. USA 96, 5263–5267 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ernst, A. et al. Neurogenesis in the striatum of the adult human brain. Cell 156, 1072–1083 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Spalding, K. L. et al. Dynamics of hippocampal neurogenesis in adult humans. Cell 153, 1219–1227 (2013). This study demonstrates for the first time that neurogenesis continues to occur in the hippocampus of adult humans.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Ryu, J. R. et al. Control of adult neurogenesis by programmed cell death in the mammalian brain. Mol. Brain 9, 43 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ernst, A. & Frisen, J. Adult neurogenesis in humans — common and unique traits in mammals. PLoS Biol. 13, e1002045 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Emsley, J. G., Mitchell, B. D., Kempermann, G. & Macklis, J. D. Adult neurogenesis and repair of the adult CNS with neural progenitors, precursors, and stem cells. Prog. Neurobiol. 75, 321–341 (2005).

    Article  CAS  PubMed  Google Scholar 

  30. Kokoeva, M. V., Yin, H. & Flier, J. S. Neurogenesis in the hypothalamus of adult mice: potential role in energy balance. Science 310, 679–683 (2005).

    Article  CAS  PubMed  Google Scholar 

  31. Contestabile, A. et al. Lithium rescues synaptic plasticity and memory in Down syndrome mice. J. Clin. Invest. 123, 348–361 (2013).

    Article  CAS  PubMed  Google Scholar 

  32. Herskovits, A. Z. & Guarente, L. SIRT1 in neurodevelopment and brain senescence. Neuron 81, 471–483 (2014). This review article summarizes important recent findings regardingsirtuins in the brain function.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gao, J. et al. A novel pathway regulates memory and plasticity via SIRT1 and miR-134. Nature 466, 1105–1109 (2010). This study demonstrates the role of SIRT1 in the regulation of synaptic plasticity and memory function through CREB.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Michan, S. et al. SIRT1 is essential for normal cognitive function and synaptic plasticity. J. Neurosci. 30, 9695–9707 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Jiang, M. et al. Neuroprotective role of Sirt1 in mammalian models of Huntington's disease through activation of multiple Sirt1 targets. Nat. Med. 18, 153–158 (2012).

    Article  CAS  Google Scholar 

  36. Jeong, H. et al. Sirt1 mediates neuroprotection from mutant huntingtin by activation of the TORC1 and CREB transcriptional pathway. Nat. Med. 18, 159–165 (2012).

    Article  CAS  Google Scholar 

  37. Corpas, R. et al. SIRT1 overexpression in mouse hippocampus induces cognitive enhancement through proteostatic and neurotrophic mechanisms. Mol. Neurobiol. http://dx.doi.org/10.1007/s12035-016-0087-9 (2016).

  38. CONVERGE consortium. Sparse whole-genome sequencing identifies two loci for major depressive disorder. Nature 523, 588–591 (2015). This is the first GWAS study for major depressive disorder to demonstrate that one of the replicated risk loci of major depressive disorder is close to the Sirt1 gene.

  39. Kishi, T. et al. SIRT1 gene is associated with major depressive disorder in the Japanese population. J. Affect. Disord. 126, 167–173 (2010).

    Article  CAS  PubMed  Google Scholar 

  40. Abe, N. et al. Altered sirtuin deacetylase gene expression in patients with a mood disorder. J. Psychiatr. Res. 45, 1106–1112 (2011).

    Article  PubMed  Google Scholar 

  41. Liu, R. et al. SIRT2 is involved in the modulation of depressive behaviors. Sci. Rep. 5, 8415 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Libert, S. et al. SIRT1 activates MAO-A in the brain to mediate anxiety and exploratory drive. Cell 147, 1459–1472 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Erikson, G. A. et al. Whole-genome sequencing of a healthy aging cohort. Cell 165, 1002–1011 (2016). This recent GWAS study of a healthy ageing cohort suggests that healthy ageing is linked to protection against cognitive decline.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Sakamoto, J., Miura, T., Shimamoto, K. & Horio, Y. Predominant expression of Sir2α, an NAD-dependent histone deacetylase, in the embryonic mouse heart and brain. FEBS Lett. 556, 281–286 (2004).

    Article  CAS  PubMed  Google Scholar 

  45. Prozorovski, T. et al. Sirt1 contributes critically to the redox-dependent fate of neural progenitors. Nat. Cell Biol. 10, 385–394 (2008).

    Article  CAS  PubMed  Google Scholar 

  46. Ichi, S. et al. Role of Pax3 acetylation in the regulation of Hes1 and Neurog2. Mol. Biol. Cell 22, 503–512 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Tiberi, L. et al. BCL6 controls neurogenesis through Sirt1-dependent epigenetic repression of selective Notch targets. Nat. Neurosci. 15, 1627–1635 (2012).

    Article  CAS  PubMed  Google Scholar 

  48. Saharan, S., Jhaveri, D. J. & Bartlett, P. F. SIRT1 regulates the neurogenic potential of neural precursors in the adult subventricular zone and hippocampus. J. Neurosci. Res. 91, 642–659 (2013).

    Article  CAS  PubMed  Google Scholar 

  49. Hisahara, S. et al. Histone deacetylase SIRT1 modulates neuronal differentiation by its nuclear translocation. Proc. Natl Acad. Sci. USA 105, 15599–15604 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Ma, C. Y. et al. SIRT1 suppresses self-renewal of adult hippocampal neural stem cells. Development 141, 4697–4709 (2014).

    Article  CAS  PubMed  Google Scholar 

  51. Stein, L. R. & Imai, S. Specific ablation of Nampt in adult neural stem cells recapitulates their functional defects during aging. EMBO J. 33, 1321–1340 (2014). This study demonstrates that NAMPT-mediated NAD+ biosynthesis is crucial for NSC self-renewal and differentiation into oligodendrocytes and that both SIRT1 and SIRT2 have redundant functions in NAMPT downstream signalling.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. van Praag, H., Shubert, T., Zhao, C. & Gage, F. H. Exercise enhances learning and hippocampal neurogenesis in aged mice. J. Neurosci. 25, 8680–8685 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Shimazu, K. et al. NT-3 facilitates hippocampal plasticity and learning and memory by regulating neurogenesis. Learn. Mem. 13, 307–315 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhao, X. et al. Mice lacking methyl-CpG binding protein 1 have deficits in adult neurogenesis and hippocampal function. Proc. Natl Acad. Sci. USA 100, 6777–6782 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Frankland, P. W. & Miller, F. D. Regenerating your senses: multiple roles for neurogenesis in the adult brain. Nat. Neurosci. 11, 1124–1126 (2008).

    Article  CAS  PubMed  Google Scholar 

  56. Braidy, N. et al. Differential expression of sirtuins in the aging rat brain. Front. Cell. Neurosci. 9, 167 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. National Association of Chronic Disease Directors. The state of mental health and aging in America: issue brief 1: what do the data tell us? (CDC, 2008).

  58. Dell'Osso, B., Buoli, M., Baldwin, D. S. & Altamura, A. C. Serotonin norepinephrine reuptake inhibitors (SNRIs) in anxiety disorders: a comprehensive review of their clinical efficacy. Hum. Psychopharmacol. 25, 17–29 (2010).

    Article  CAS  PubMed  Google Scholar 

  59. File, S. E., Gonzalez, L. E. & Andrews, N. Comparative study of pre- and postsynaptic 5-HT1A receptor modulation of anxiety in two ethological animal tests. J. Neurosci. 16, 4810–4815 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Chung, S. et al. Electroconvulsive shock increases SIRT1 immunoreactivity in the mouse hippocampus and hypothalamus. J. ECT 29, 93–100 (2013).

    Article  CAS  PubMed  Google Scholar 

  61. Preskorn, S. H. Pharmacogenomics, informatics, and individual drug therapy in psychiatry: past, present and future. J. Psychopharmacol. 20, 85–94 (2006).

    Article  PubMed  Google Scholar 

  62. Tran, L., Schulkin, J., Ligon, C. O. & Greenwood-Van Meerveld, B. Epigenetic modulation of chronic anxiety and pain by histone deacetylation. Mol. Psychiatry 20, 1219–1231 (2015).

    Article  CAS  PubMed  Google Scholar 

  63. Tsukahara, S., Tanaka, S., Ishida, K., Hoshi, N. & Kitagawa, H. Age-related change and its sex differences in histoarchitecture of the hypothalamic suprachiasmatic nucleus of F344/N rats. Exp. Gerontol. 40, 147–155 (2005).

    Article  PubMed  Google Scholar 

  64. Roberts, D. E., Killiany, R. J. & Rosene, D. L. Neuron numbers in the hypothalamus of the normal aging rhesus monkey: stability across the adult lifespan and between the sexes. J. Comp. Neurol. 520, 1181–1197 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Nygard, M., Hill, R. H., Wikstrom, M. A. & Kristensson, K. Age-related changes in electrophysiological properties of the mouse suprachiasmatic nucleus in vitro. Brain Res. Bull. 65, 149–154 (2005).

    Article  PubMed  Google Scholar 

  66. Farajnia, S., Meijer, J. H. & Michel, S. Age-related changes in large-conductance calcium-activated potassium channels in mammalian circadian clock neurons. Neurobiol. Aging 36, 2176–2183 (2015).

    Article  CAS  PubMed  Google Scholar 

  67. Chang, H. C. & Guarente, L. SIRT1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell 153, 1448–1460 (2013). This study demonstrates that decline in SIRT1 in the SCN causes circadian dysfunction in aged mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Libert, S., Bonkowski, M. S., Pointer, K., Pletcher, S. D. & Guarente, L. Deviation of innate circadian period from 24 h reduces longevity in mice. Aging Cell 11, 794–800 (2012).

    Article  CAS  PubMed  Google Scholar 

  69. Hasan, S., Dauvilliers, Y., Mongrain, V., Franken, P. & Tafti, M. Age-related changes in sleep in inbred mice are genotype dependent. Neurobiol. Aging 33, 195.e13–195.e26 (2012).

    Article  Google Scholar 

  70. Landolt, H. P., Dijk, D. J., Achermann, P. & Borbely, A. A. Effect of age on the sleep EEG: slow-wave activity and spindle frequency activity in young and middle-aged men. Brain Res. 738, 205–212 (1996).

    Article  CAS  PubMed  Google Scholar 

  71. Dew, M. A. et al. Healthy older adults' sleep predicts all-cause mortality at 4 to 19 years of follow-up. Psychosom. Med. 65, 63–73 (2003).

    Article  PubMed  Google Scholar 

  72. Tasali, E., Leproult, R., Ehrmann, D. A. & Van Cauter, E. Slow-wave sleep and the risk of type 2 diabetes in humans. Proc. Natl Acad. Sci. USA 105, 1044–1049 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  73. Nohara, K., Yoo, S. H. & Chen, Z. J. Manipulating the circadian and sleep cycles to protect against metabolic disease. Front. Endocrinol. (Lausanne) 6, 35 (2015).

    Article  Google Scholar 

  74. Panossian, L. et al. SIRT1 regulation of wakefulness and senescence-like phenotype in wake neurons. J. Neurosci. 31, 4025–4036 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Zhang, J. et al. Extended wakefulness: compromised metabolics in and degeneration of locus ceruleus neurons. J. Neurosci. 34, 4418–4431 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Gutierrez, E. G., Banks, W. A. & Kastin, A. J. Murine tumor necrosis factor alpha is transported from blood to brain in the mouse. J. Neuroimmunol. 47, 169–176 (1993).

    Article  CAS  PubMed  Google Scholar 

  77. Banks, W. A. & Kastin, A. J. Reversible association of the cytokines MIP-1α and MIP-1ß with the endothelia of the blood-brain barrier. Neurosci. Lett. 205, 202–206 (1996).

    Article  CAS  PubMed  Google Scholar 

  78. Pan, W., Banks, W. A. & Kastin, A. J. Permeability of the blood-brain and blood-spinal cord barriers to interferons. J. Neuroimmunol. 76, 105–111 (1997).

    Article  CAS  PubMed  Google Scholar 

  79. Pan, W., Banks, W. A., Fasold, M. B., Bluth, J. & Kastin, A. J. Transport of brain-derived neurotrophic factor across the blood-brain barrier. Neuropharmacology 37, 1553–1561 (1998).

    Article  CAS  PubMed  Google Scholar 

  80. Poduslo, J. F. & Curran, G. L. Permeability at the blood-brain and blood-nerve barriers of the neurotrophic factors: NGF, CNTF, NT-3, BDNF. Brain Res. Mol. Brain Res. 36, 280–286 (1996).

    Article  CAS  PubMed  Google Scholar 

  81. Spranger, J. et al. Adiponectin does not cross the blood-brain barrier but modifies cytokine expression of brain endothelial cells. Diabetes 55, 141–147 (2006).

    Article  CAS  PubMed  Google Scholar 

  82. Pan, W. & Kastin, A. J. Changing the chemokine gradient: CINC1 crosses the blood-brain barrier. J. Neuroimmunol. 115, 64–70 (2001).

    Article  CAS  PubMed  Google Scholar 

  83. Banks, W. A. Cytokines and the Blood-Brain Barrier (eds Seigel, A. & Zalcman, S. S.) 3–17 (Springer, 2009).

    Google Scholar 

  84. Xie, Z., Morgan, T. E., Rozovsky, I. & Finch, C. E. Aging and glial responses to lipopolysaccharide in vitro: greater induction of IL-1 and IL-6, but smaller induction of neurotoxicity. Exp. Neurol. 182, 135–141 (2003).

    Article  CAS  PubMed  Google Scholar 

  85. Maher, F. O., Martin, D. S. & Lynch, M. A. Increased IL-1ß in cortex of aged rats is accompanied by downregulation of ERK and PI-3 kinase. Neurobiol. Aging 25, 795–806 (2004).

    Article  CAS  PubMed  Google Scholar 

  86. Godbout, J. P. et al. Exaggerated neuroinflammation and sickness behavior in aged mice following activation of the peripheral innate immune system. FASEB J. 19, 1329–1331 (2005).

    Article  CAS  PubMed  Google Scholar 

  87. Maher, F. O., Nolan, Y. & Lynch, M. A. Downregulation of IL-4-induced signalling in hippocampus contributes to deficits in LTP in the aged rat. Neurobiol. Aging 26, 717–728 (2005).

    Article  CAS  PubMed  Google Scholar 

  88. Ye, S. M. & Johnson, R. W. An age-related decline in interleukin-10 may contribute to the increased expression of interleukin-6 in brain of aged mice. Neuroimmunomodulation 9, 183–192 (2001).

    Article  CAS  PubMed  Google Scholar 

  89. Lee, P. et al. Effects of aging on blood brain barrier and matrix metalloproteases following controlled cortical impact in mice. Exp. Neurol. 234, 50–61 (2012).

    Article  CAS  PubMed  Google Scholar 

  90. Elahy, M. et al. Blood-brain barrier dysfunction developed during normal aging is associated with inflammation and loss of tight junctions but not with leukocyte recruitment. Immun. Ageing 12, 2 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Henry, C. J., Huang, Y., Wynne, A. M. & Godbout, J. P. Peripheral lipopolysaccharide (LPS) challenge promotes microglial hyperactivity in aged mice that is associated with exaggerated induction of both pro-inflammatory IL-1ß and anti-inflammatory IL-10 cytokines. Brain Behav. Immun. 23, 309–317 (2009).

    Article  CAS  PubMed  Google Scholar 

  92. Wong, A. M. et al. Macrosialin increases during normal brain aging are attenuated by caloric restriction. Neurosci. Lett. 390, 76–80 (2005).

    Article  CAS  PubMed  Google Scholar 

  93. Perry, V. H., Matyszak, M. K. & Fearn, S. Altered antigen expression of microglia in the aged rodent CNS. Glia 7, 60–67 (1993).

    Article  CAS  PubMed  Google Scholar 

  94. Stichel, C. C. & Luebbert, H. Inflammatory processes in the aging mouse brain: participation of dendritic cells and T-cells. Neurobiol. Aging 28, 1507–1521 (2007).

    Article  CAS  PubMed  Google Scholar 

  95. Letiembre, M. et al. Innate immune receptor expression in normal brain aging. Neuroscience 146, 248–254 (2007).

    Article  CAS  PubMed  Google Scholar 

  96. Griffin, R. et al. The age-related attenuation in long-term potentiation is associated with microglial activation. J. Neurochem. 99, 1263–1272 (2006).

    Article  CAS  PubMed  Google Scholar 

  97. Griciuc, A. et al. Alzheimer's disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 78, 631–643 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Jonsson, T. et al. Variant of TREM2 associated with the risk of Alzheimer's disease. N. Engl. J. Med. 368, 107–116 (2013).

    Article  CAS  PubMed  Google Scholar 

  99. Nimmerjahn, A., Kirchhoff, F. & Helmchen, F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308, 1314–1318 (2005).

    Article  CAS  PubMed  Google Scholar 

  100. Chen, J. et al. SIRT1 protects against microglia-dependent amyloid-ß toxicity through inhibiting NF-κB signaling. J. Biol. Chem. 280, 40364–40374 (2005).

    Article  CAS  PubMed  Google Scholar 

  101. Li, L. et al. Overexpression of SIRT1 induced by resveratrol and inhibitor of miR-204 suppresses activation and proliferation of microglia. J. Mol. Neurosci. 56, 858–867 (2015).

    Article  CAS  PubMed  Google Scholar 

  102. Cho, S. H. et al. SIRT1 deficiency in microglia contributes to cognitive decline in aging and neurodegeneration via epigenetic regulation of IL-1ß. J. Neurosci. 35, 807–818 (2015). This study shows that reduction in microglial SIRT1 causes age-associated memory deficits via IL-1 β upregulation in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Pais, T. F. et al. The NAD-dependent deacetylase sirtuin 2 is a suppressor of microglial activation and brain inflammation. EMBO J. 32, 2603–2616 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Mohammed, C. P. et al. miR-204 downregulates EphB2 in aging mouse hippocampal neurons. Aging Cell 15, 380–388 (2016).

    Article  CAS  PubMed  Google Scholar 

  105. Qin, W. et al. Neuronal SIRT1 activation as a novel mechanism underlying the prevention of Alzheimer disease amyloid neuropathology by calorie restriction. J. Biol. Chem. 281, 21745–21754 (2006).

    Article  CAS  PubMed  Google Scholar 

  106. Kim, D. et al. SIRT1 deacetylase protects against neurodegeneration in models for Alzheimer's disease and amyotrophic lateral sclerosis. EMBO J. 26, 3169–3179 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Min, S. W. et al. Acetylation of tau inhibits its degradation and contributes to tauopathy. Neuron 67, 953–966 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Bonda, D. J. et al. The sirtuin pathway in ageing and Alzheimer disease: mechanistic and therapeutic considerations. Lancet Neurol. 10, 275–279 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Guo, Y. J. et al. Resveratrol alleviates MPTP-induced motor impairments and pathological changes by autophagic degradation of α-synuclein via SIRT1-deacetylated LC3. Mol. Nutr. Food Res. 60, 2161–2175 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Ferretta, A. et al. Effect of resveratrol on mitochondrial function: implications in parkin-associated familiar Parkinson's disease. Biochim. Biophys. Acta 1842, 902–915 (2014).

    Article  CAS  PubMed  Google Scholar 

  111. Parker, J. A. et al. Resveratrol rescues mutant polyglutamine cytotoxicity in nematode and mammalian neurons. Nat. Genet. 37, 349–350 (2005).

    Article  CAS  PubMed  Google Scholar 

  112. Parker, J. A. et al. Integration of ß-catenin, sirtuin, and FOXO signaling protects from mutant huntingtin toxicity. J. Neurosci. 32, 12630–12640 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Wareski, P. et al. PGC-1α and PGC-1β regulate mitochondrial density in neurons. J. Biol. Chem. 284, 21379–21385 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Han, S., Choi, J. R., Soon Shin, K. & Kang, S. J. Resveratrol upregulated heat shock proteins and extended the survival of G93A-SOD1 mice. Brain Res. 1483, 112–117 (2012).

    Article  CAS  PubMed  Google Scholar 

  115. Yanez, M. et al. CSF from amyotrophic lateral sclerosis patients produces glutamate independent death of rat motor brain cortical neurons: protection by resveratrol but not riluzole. Brain Res. 1423, 77–86 (2011).

    Article  CAS  PubMed  Google Scholar 

  116. Outeiro, T. F. et al. Sirtuin 2 inhibitors rescue α-synuclein-mediated toxicity in models of Parkinson's disease. Science 317, 516–519 (2007).

    Article  CAS  PubMed  Google Scholar 

  117. Bobrowska, A., Donmez, G., Weiss, A., Guarente, L. & Bates, G. SIRT2 ablation has no effect on tubulin acetylation in brain, cholesterol biosynthesis or the progression of Huntington's disease phenotypes in vivo. PLoS ONE 7, e34805 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Fu, J. et al. trans-(–)-epsilon-Viniferin increases mitochondrial sirtuin 3 (SIRT3), activates AMP-activated protein kinase (AMPK), and protects cells in models of Huntington Disease. J. Biol. Chem. 287, 24460–24472 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Luthi-Carter, R. et al. SIRT2 inhibition achieves neuroprotection by decreasing sterol biosynthesis. Proc. Natl Acad. Sci. USA 107, 7927–7932 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  120. Dai, S. H. et al. Sirt3 protects cortical neurons against oxidative stress via regulating mitochondrial Ca2+ and mitochondrial biogenesis. Int. J. Mol. Sci. 15, 14591–14609 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Someya, S., Yamasoba, T., Weindruch, R., Prolla, T. A. & Tanokura, M. Caloric restriction suppresses apoptotic cell death in the mammalian cochlea and leads to prevention of presbycusis. Neurobiol. Aging 28, 1613–1622 (2007).

    Article  PubMed  Google Scholar 

  122. Someya, S. et al. Sirt3 mediates reduction of oxidative damage and prevention of age-related hearing loss under caloric restriction. Cell 143, 802–812 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Kim, S. H., Lu, H. F. & Alano, C. C. Neuronal Sirt3 protects against excitotoxic injury in mouse cortical neuron culture. PLoS ONE 6, e14731 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Song, W., Song, Y., Kincaid, B., Bossy, B. & Bossy-Wetzel, E. Mutant SOD1G93A triggers mitochondrial fragmentation in spinal cord motor neurons: neuroprotection by SIRT3 and PGC-1α. Neurobiol. Dis. 51, 72–81 (2013).

    Article  CAS  PubMed  Google Scholar 

  125. Herskovits, A. Z. & Guarente, L. Sirtuin deacetylases in neurodegenerative diseases of aging. Cell Res. 23, 746–758 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Bishop, N. A. & Guarente, L. Two neurons mediate diet-restriction-induced longevity in C. elegans. Nature 447, 545–549 (2007).

    Article  CAS  PubMed  Google Scholar 

  127. Zhang, G. et al. Hypothalamic programming of systemic ageing involving IKK-ß, NF-κB and GnRH. Nature 497, 211–216 (2013). This study demonstrates that NF- κ B signalling in the hypothalamus has a crucial role in the regulation of mammalian ageing and longevity.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Bluher, M., Kahn, B. B. & Kahn, C. R. Extended longevity in mice lacking the insulin receptor in adipose tissue. Science 299, 572–574 (2003).

    Article  CAS  PubMed  Google Scholar 

  129. Revollo, J. R., Grimm, A. A. & Imai, S. The NAD biosynthesis pathway mediated by nicotinamide phosphoribosyltransferase regulates Sir2 activity in mammalian cells. J. Biol. Chem. 279, 50754–50763 (2004).

    Article  CAS  PubMed  Google Scholar 

  130. Revollo, J. R. et al. Nampt/PBEF/visfatin regulates insulin secretion in beta cells as a systemic NAD biosynthetic enzyme. Cell Metab. 6, 363–375 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Yoon, M. J. et al. SIRT1-mediated eNAMPT secretion from adipose tissue regulates hypothalamic NAD+ and function in mice. Cell Metab. 21, 706–717 (2015). This study demonstrates that eNAMPT secretion is regulated by SIRT1 in white adipose tissue and that eNAMPT controls hypothalamic NAD+ signallingand function.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Zhao, Y. et al. Extracellular visfatin has nicotinamide phosphoribosyltransferase enzymatic activity and is neuroprotective against ischemic injury. CNS Neurosci. Ther. 20, 539–547 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Jing, Z. et al. Neuronal NAMPT is released after cerebral ischemia and protects against white matter injury. J. Cereb. Blood Flow Metab. 34, 1613–1621 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Erfani, S. et al. Nampt/PBEF/visfatin exerts neuroprotective effects against ischemia/reperfusion injury via modulation of Bax/Bcl-2 ratio and prevention of caspase-3 activation. J. Mol. Neurosci. 56, 237–243 (2015).

    Article  CAS  PubMed  Google Scholar 

  135. Erfani, S. et al. Visfatin reduces hippocampal CA1 cells death and improves learning and memory deficits after transient global ischemia/reperfusion. Neuropeptides 49, 63–68 (2015).

    Article  CAS  PubMed  Google Scholar 

  136. Ramsey, K. M., Mills, K. F., 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 7, 78–88 (2008).

    Article  CAS  PubMed  Google Scholar 

  137. Gomes, A. P. et al. Declining NAD+ induces a pseudohypoxic state disrupting nuclear-mitochondrial communication during aging. Cell 155, 1624–1638 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. de Picciotto, N. E. et al. Nicotinamide mononucleotide supplementation reverses vascular dysfunction and oxidative stress with aging in mice. Aging Cell 15, 522–530 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Mills, K. F. et al. Long-term administration of nicotinamide mononucleotide mitigates age-associated physiological decline in mice. Cell Metab. 24, 795–806 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Yoshino, J., Mills, K. F., Yoon, M. J. & Imai, S. Nicotinamide mononucleotide, a key NAD+ intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 14, 528–536 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Zhao, Y. et al. Regenerative neurogenesis after ischemic stroke promoted by nicotinamide phosphoribosyltransferase-nicotinamide adenine dinucleotide cascade. Stroke 46, 1966–1974 (2015).

    Article  CAS  PubMed  Google Scholar 

  142. Wang, X., Hu, X., Yang, Y., Takata, T. & Sakurai, T. Nicotinamide mononucleotide protects against ß-amyloid oligomer- induced cognitive impairment and neuronal death. Brain Res. 1643, 1–9 (2016).

    Article  CAS  PubMed  Google Scholar 

  143. Long, A. N. et al. Effect of nicotinamide mononucleotide on brain mitochondrial respiratory deficits in an Alzheimer's disease-relevant murine model. BMC Neurol. 15, 19 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Canto, C. et al. The NAD+ precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity. Cell Metab. 15, 838–847 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Gong, B. et al. Nicotinamide riboside restores cognition through an upregulation of proliferator-activated receptor-γ coactivator 1α regulated ß-secretase 1 degradation and mitochondrial gene expression in Alzheimer's mouse models. Neurobiol. Aging 34, 1581–1588 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Zhang, H. et al. NAD+ repletion improves mitochondrial and stem cell function and enhances life span in mice. Science 352, 1436–1443 (2016). This study demonstrates that nicotinamide riboside treatment improves muscle stem cell function in aged mice and extends lifespan when the treatment starts at 2 years of age.

    Article  CAS  PubMed  Google Scholar 

  147. Scudellari, M. Ageing research: blood to blood. Nature 517, 426–429 (2015).

    Article  CAS  PubMed  Google Scholar 

  148. Villeda, S. A. et al. The ageing systemic milieu negatively regulates neurogenesis and cognitive function. Nature 477, 90–94 (2011). This study demonstrates that CCL11, a circulating blood-borne factor identified by heterochronic parabiosis, affects adult neurogenesis and impairs cognitive function in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Smith, L. K. et al. β2-microglobulin is a systemic pro-aging factor that impairs cognitive function and neurogenesis. Nat. Med. 21, 932–937 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Ruckh, J. M. et al. Rejuvenation of regeneration in the aging central nervous system. Cell Stem Cell 10, 96–103 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Itoh, N. FGF21 as a hepatokine, adipokine, and myokine in metabolism and diseases. Front. Endocrinol. (Lausanne) 5, 107 (2014).

    Article  Google Scholar 

  152. Zhang, Y. et al. The starvation hormone, fibroblast growth factor-21, extends lifespan in mice. eLife 1, e00065 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Mendelsohn, A. R. & Larrick, J. W. Fibroblast growth factor-21 is a promising dietary restriction mimetic. Rejuvenation Res. 15, 624–628 (2012).

    Article  CAS  PubMed  Google Scholar 

  154. Kharitonenkov, A. et al. FGF-21/FGF-21 receptor interaction and activation is determined by βKlotho. J. Cell. Physiol. 215, 1–7 (2008).

    Article  CAS  PubMed  Google Scholar 

  155. Owen, B. M. et al. FGF21 contributes to neuroendocrine control of female reproduction. Nat. Med. 19, 1153–1156 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Bookout, A. L. et al. FGF21 regulates metabolism and circadian behavior by acting on the nervous system. Nat. Med. 19, 1147–1152 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Yu, Y. et al. Fibroblast growth factor 21 protects mouse brain against D-galactose induced aging via suppression of oxidative stress response and advanced glycation end products formation. Pharmacol. Biochem. Behav. 133, 122–131 (2015).

    Article  CAS  PubMed  Google Scholar 

  158. Katsimpardi, L. et al. Vascular and neurogenic rejuvenation of the aging mouse brain by young systemic factors. Science 344, 630–634 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Walker, R. G. et al. Biochemistry and biology of GDF11 and myostatin: similarities, differences, and questions for future investigation. Circ. Res. 118, 1125–1141 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Poggioli, T. et al. Circulating growth differentiation factor 11/8 levels decline with age. Circ. Res. 118, 29–37 (2016).

    Article  CAS  PubMed  Google Scholar 

  161. Tsai, S. et al. Muscle-specific 4E-BP1 signaling activation improves metabolic parameters during aging and obesity. J. Clin. Invest. 125, 2952–2964 (2015). This study demonstrates that skeletal muscle-specific activation of 4EBP1, one of the key downstream targets of the mTOR pathway, prevents age-induced and diet-induced insulin resistance and a decline in metabolic rate.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Pedersen, B. K. & Febbraio, M. A. Muscle as an endocrine organ: focus on muscle-derived interleukin-6. Physiol. Rev. 88, 1379–1406 (2008).

    Article  CAS  PubMed  Google Scholar 

  163. Banks, W. A., Kastin, A. J. & Gutierrez, E. G. Penetration of interleukin-6 across the murine blood-brain barrier. Neurosci. Lett. 179, 53–56 (1994).

    Article  CAS  PubMed  Google Scholar 

  164. Singh, T. & Newman, A. B. Inflammatory markers in population studies of aging. Ageing Res. Rev. 10, 319–329 (2011).

    Article  CAS  PubMed  Google Scholar 

  165. Wrann, C. D. et al. Exercise induces hippocampal BDNF through a PGC-1α/FNDC5 pathway. Cell Metab. 18, 649–659 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Xu, H. & Heilshorn, S. C. Microfluidic investigation of BDNF-enhanced neural stem cell chemotaxis in CXCL12 gradients. Small 9, 585–595 (2013).

    Article  CAS  PubMed  Google Scholar 

  167. Li, X., Khanna, A., Li, N. & Wang, E. Circulatory miR34a as an RNAbased, noninvasive biomarker for brain aging. Aging (Albany NY) 3, 985–1002 (2011).

    Article  CAS  Google Scholar 

  168. Dhahbi, J. M. et al. Deep sequencing identifies circulating mouse miRNAs that are functionally implicated in manifestations of aging and responsive to calorie restriction. Aging (Albany NY) 5, 130–141 (2013).

    Article  CAS  Google Scholar 

  169. Noren Hooten, N. et al. Age-related changes in microRNA levels in serum. Aging (Albany NY) 5, 725–740 (2013).

    Article  Google Scholar 

  170. Olivieri, F. et al. Age-related differences in the expression of circulating microRNAs: miR-21 as a new circulating marker of inflammaging. Mech. Ageing Dev. 133, 675–685 (2012).

    Article  CAS  PubMed  Google Scholar 

  171. Scheubel, R. J. et al. Age-dependent depression in circulating endothelial progenitor cells in patients undergoing coronary artery bypass grafting. J. Am. Coll. Cardiol. 42, 2073–2080 (2003).

    Article  PubMed  Google Scholar 

  172. Zierer, J., Menni, C., Kastenmuller, G. & Spector, T. D. Integration of 'omics' data in aging research: from biomarkers to systems biology. Aging Cell 14, 933–944 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Valdes, A. M., Glass, D. & Spector, T. D. Omics technologies and the study of human ageing. Nat. Rev. Genet. 14, 601–607 (2013).

    Article  CAS  PubMed  Google Scholar 

  174. Neufer, P. D. et al. Understanding the cellular and molecular mechanisms of physical activity-induced health benefits. Cell Metab. 22, 4–11 (2015).

    Article  CAS  PubMed  Google Scholar 

  175. Profant, O. et al. Functional changes in the human auditory cortex in ageing. PLoS ONE 10, e0116692 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Perez, P. & Bao, J. Why do hair cells and spiral ganglion neurons in the cochlea die during aging? Aging Dis. 2, 231–241 (2011).

    PubMed  PubMed Central  Google Scholar 

  177. Sweet, R. J., Price, J. M. & Henry, K. R. Dietary restriction and presbyacusis: periods of restriction and auditory threshold losses in the CBA/J mouse. Audiology 27, 305–312 (1988).

    Article  CAS  PubMed  Google Scholar 

  178. Quan, Y. et al. Adjudin protects rodent cochlear hair cells against gentamicin ototoxicity via the SIRT3-ROS pathway. Sci. Rep. 5, 8181 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Xiong, H. et al. SIRT1 expression in the cochlea and auditory cortex of a mouse model of age-related hearing loss. Exp. Gerontol. 51, 8–14 (2014).

    Article  CAS  PubMed  Google Scholar 

  180. Xiong, H. et al. Activation of miR-34a/SIRT1/p53 signaling contributes to cochlear hair cell apoptosis: implications for age-related hearing loss. Neurobiol. Aging 36, 1692–1701 (2015).

    Article  CAS  PubMed  Google Scholar 

  181. Han, C. et al. Sirt1 deficiency protects cochlear cells and delays the early onset of age-related hearing loss in C57BL/6 mice. Neurobiol. Aging 43, 58–71 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Miskin, R. & Masos, T. Transgenic mice overexpressing urokinase-type plasminogen activator in the brain exhibit reduced food consumption, body weight and size, and increased longevity. J. Gerontol. A Biol. Sci. Med. Sci. 52, B118–B124 (1997).

    Article  CAS  PubMed  Google Scholar 

  183. Conti, B. et al. Transgenic mice with a reduced core body temperature have an increased life span. Science 314, 825–828 (2006).

    Article  CAS  PubMed  Google Scholar 

  184. Taguchi, A., Wartschow, L. M. & White, M. F. Brain IRS2 signaling coordinates life span and nutrient homeostasis. Science 317, 369–372 (2007).

    Article  CAS  PubMed  Google Scholar 

  185. Kappeler, L. et al. Brain IGF-1 receptors control mammalian growth and lifespan through a neuroendocrine mechanism. PLoS Biol. 6, e254 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Tang, Y., Purkayastha, S. & Cai, D. Hypothalamic microinflammation: a common basis of metabolic syndrome and aging. Trends Neurosci. 38, 36–44 (2015).

    Article  CAS  PubMed  Google Scholar 

  187. Brack, A. S. et al. Increased Wnt signaling during aging alters muscle stem cell fate and increases fibrosis. Science 317, 807–810 (2007).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors apologize to those whose work is not cited owing to space limitations. The authors thank members of the Imai laboratory, Cynthia S. Brace in particular, for stimulating discussions and critical reading of the manuscript. This work was supported by grants from the National Institute on Aging (AG037457 and AG047902) to S.I. and from the American Sleep Medicine Foundation to A.S., and by the Glenn Foundation for Medical Research and grants from the National Institutes of Health to L.G.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Leonard Guarente.

Ethics declarations

Competing interests

L.G. is a founder of Elysium Health and consults for GSK, Chronos (Oxford) and Segterra. A.S. and S.I. declare no competing interests.

PowerPoint slides

Glossary

Myokines

Biologically active peptides produced and secreted by muscle cells that mediate a range of biological effects in both original and remote tissues and organs.

Hepatokines

Biologically active peptides produced and secreted by hepatocytes that mediate a range of biological effects in both original and remote tissues and organs.

Adipokines

Biologically active peptides produced and secreted by adipocytes that mediate a range of biological effects in both original and remote tissues and organs.

Cytokines

General terms for small-size cell signalling peptides such as lymphokines, monokines, chemokines and interleukins (cytokines produced by lymphocytes, monocytes, chemocytes and leukocytes, respectively).

Brain atrophy

Shrinkage of the brain described as a loss of neurons and connections between neurons.

Circadian clock

Central mechanisms that drive circadian rhythm.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Satoh, A., Imai, Si. & Guarente, L. The brain, sirtuins, and ageing. Nat Rev Neurosci 18, 362–374 (2017). https://doi.org/10.1038/nrn.2017.42

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrn.2017.42

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing