Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Nov 10.
Published in final edited form as: Nat Cell Biol. 2019 Jan 2;21(1):32–43. doi: 10.1038/s41556-018-0206-0

Turning back time with emerging rejuvenation strategies

Salah Mahmoudi 1,4, Lucy Xu 1,2,4, Anne Brunet 1,3,*
PMCID: PMC7653017  NIHMSID: NIHMS1638129  PMID: 30602763

Abstract

Ageing is associated with the functional decline of all tissues and a striking increase in many diseases. Although ageing has long been considered a one-way street, strategies to delay and potentially even reverse the ageing process have recently been developed. Here, we review four emerging rejuvenation strategies—systemic factors, metabolic manipulations, senescent cell ablation and cellular reprogramming—and discuss their mechanisms of action, cellular targets, potential trade-offs and application to human ageing.


Ageing represents a major risk factor for many chronic conditions, including cardiovascular disease, diabetes, cancer, arthritis and frailty1,2. Once considered irreversible, ageing is in fact remarkably malleable. Indeed, inhibition of high-nutrient-sensing pathways (for example, the insulin–insulin-like growth factor (IGF) and mechanistic target of rapamycin (mTOR) pathways) and activation of low-nutrient-sensing proteins (for example, 5′ AMP-activated protein kinase (AMPK) and sirtuins) extend lifespan in various model organisms3,4. Diet-based interventions, such as dietary restriction, and pharmacological interventions, including the mTOR inhibitor rapamycin, improve aspects of ageing even when administered late in life5-10. A key question is whether ageing of cells, tissues and organisms can be reversed or ‘rejuvenated’ rather than simply delayed.

A host of age-associated features have been identified, with a subset being potential drivers of the ageing process (extensively reviewed elsewhere1,2). At the molecular level, ageing hallmarks comprise DNA damage, epigenetic alterations, telomere attrition, protein aggregation and accumulation of aberrant mitochondria and lysosomes1,2. At the cellular and organismal level, ageing features include cellular senescence, stem cell exhaustion, deregulated nutrient sensing and chronic low-grade inflammation1,2. Various rejuvenation strategies that target these hallmarks have recently emerged and they fall into four broad categories: systemic (blood) factors, metabolic manipulations, senescent cell ablation and cellular reprogramming. Although these approaches seemingly target very different ageing features11-15, a central question is whether they share common mechanisms of action. This Review discusses these four rejuvenation strategies and how they improve health and lifespan. We also address several key questions: which hallmarks of ageing are targeted by each strategy and are there commonalities in their modes of action? Does the rejuvenating effect come with trade-offs? Ultimately, can rejuvenation strategies be used to improve human health and longevity and target age-associated diseases?

Blood factors as targets for rejuvenation

Heterochronic parabiosis studies, in which the circulatory systems of a young mouse and an aged mouse are fused, have provided compelling evidence that blood factors influence organismal ageing (Fig. 1, Table 1 and Supplementary Table 1). Heterochronic parabiosis was initially shown to revitalize muscle stem cells in naturally aged mice, reversing the age-dependent decline in stem cell activation and number and improving their age-associated differentiation bias16,17. Since then, heterochronic parabiosis has been shown to enhance muscle, liver, brain and heart function of aged mice17-24, by boosting the function of both stem and differentiated cells17-23. Sharing blood circulation with a young mouse also reduces genomic instability in the aged mouse20 and reverses age-associated gene expression signatures25. Blood factors, rather than blood cells, seem to play a major role in these rejuvenating effects24-26: direct injection of young blood plasma (devoid of cells)25 or of human umbilical cord plasma (also devoid of cells)26 into aged mice can recapitulate several aspects of heterochronic parabiosis, notably the increase in neurogenesis and improvement of cognitive functions25,26 (Table 1 and Supplementary Table 1). These observations raise the possibility that blood factors (for example, proteins, metabolites, lipids and exosomes) could be used to reverse aspects of the ageing process, perhaps even in humans.

Fig. 1 ∣. Comparison of emerging strategies for organismal rejuvenation and lifespan.

Fig. 1 ∣

A comparison of the four emerging rejuvenation strategies: blood factors, metabolic manipulation, ablation of senescent cells and cellular reprogramming. The figure depicts the features that improve when treatment in mice is initiated at midlife or later. The top panel shows organs or tissues that exhibit a rejuvenated phenotype in wild-type (WT) mice. For rapamycin, features that have been shown to improve also in young mice following treatment are indicated with an asterisk (*). The effect on lifespan, proposed primary mode (or modes) of action and possible trade-offs of these strategies are also presented. Finally, the translational potential in humans is indicated by the increasing number of plus signs (+) based on present evidence in human ageing and current feasibility. NT, not tested. Question marks indicate possible modes of action and trade-offs. Figure adapted from ref. 188.

Table 1 ∣.

Summary of studies testing rejuvenation interventions at midlife or later in naturally ageing mice

Intervention Age at
intervention
(months)
Metric output Comparison points (control) Ref.
Blood factors
Parabiosis 2–3, 19–26 Skeletal muscle (MuSC) and liver regeneration Old-old and young-young parabionts 17
Parabiosis 4–6, 24–26 Muscle regeneration (MuSC and fibrosis) Old-old and young-young parabionts 16
Parabiosis 3–4, 18–20 Neurogenesis and cognitive function Old-old and young-young parabionts 21
Parabiosis 1–2, 10–12 Spinal cord remyelination Middle-aged-middle-aged and young-young parabionts 24
Parabiosis 2, 23 Cardiac metrics Old-old and young-young parabionts 18
Parabiosis 2, 15–16 or 21 Neurogenesis and cognitive function Old-old and young-young parabionts 22
Parabiosis 2–3, 22–24 Muscle regeneration (MuSC) and function Old-old and young-young parabionts 20
Parabiosis 3, 18 Synaptic plasticity and gene expression Old-old parabionts 25
Parabiosis 3, 19 Bone regeneration Old-old and young-young parabionts 19
Parabiosis 3, 18 Neurogenesis and cognitive function Young-young parabionts 23
Young blood injection 18 Cognitive function and gene expression Old blood 25
Human plasma injection (cord, young and elderly) 8–10, 13–14 Neuronal and cognitive functions and gene expression Age-matched vehicle control, young (22 years of age) and old (66 years of age) human plasma
TIMP2 administration 18 Synaptic plasticity and cognitive functions Age-matched vehicle control 26
Oxytocin administration 2–4, 22–24 Muscle regeneration (MuSC and fibrosis) Age-matched vehicle and antagonist (only young) control 36
GDF11 administration 23–24 Cardiac metrics Age-matched vehicle control 18
GDF11 administration 21–23 Neurogenesis and cognitive function Age-matched vehicle control 22
GDF11 administration 2–3, 22–24 Muscle regeneration (MuSC) and function Age-matched vehicle control 20
GDF11 administration 23 Muscle regeneration (MuSC) Age-matched vehicle control 34
GDF11 administration 24 Cardiac metrics and function 2 months of age, 3 months of age and age-matched vehicle treated 33
GDF11 administration 2, 22 Cardiac metrics and body weight Age-matched vehicle control 32
Metabolic manipulation
Short-term dietary restriction 5–8, 28–30 Vasculature metrics Age-matched ad libitum 43
Short-term dietary restriction 2, 18 Skeletal muscle (MuSC) Age-matched ad libitum 7
Fasting-mimicking diet 16 Organ size and regeneration 16 months of age and age-matched ad libitum 5
Fasting-mimicking diet 16 Immunosenescence 4 months of age, 16 months of age and age-matched ad libitum 5
Fasting-mimicking diet 16 Cognitive function Age-matched ad libitum 5
Fasting-mimicking diet 16 Bone density 12 months of age and age-matched ad libitum 5
Fasting-mimicking diet 16 Cancer and inflammation Age-matched ad libitum 5
Ketogenic diet 12 Physiological and metabolic metrics; physical, behaviour and cognitive functions Age-matched ad libitum and low-carbohydrate non-ketogenic
Ketogenic diet 12–14 Cognitive and motor function and frailty index 12 months of age and age-matched ad libitum 8
Ketogenic diet 12–14 Cognitive and motor function 12 months of age and age-matched ad libitum 8
Rapamycin 22 Immune system (HSC and immunity) 2 months of age and age-matched vehicle control 48
Rapamycin 4, 13, 20–22 Comprehensive organismal assessment (>25 tissues) 3–6 months of age and age-matched vehicle control 49
Metformin 12 Serum biomarkers Age-matched ad libitum and dietary restricted 41
Metformin 12 Physical performance Age-matched ad libitum 41
Metformin 12 Liver, muscle and gene expression Age-matched ad libitum and dietary restricted 41
Resveratrol 12 Physiological metrics and gene expression Age-matched untreated controls 54
Resveratrol 18 Renal function and histology Age-matched untreated control 53
Ablation of senescent cells
Ablation of p16-positive cells 18 Adipose tissue metrics Age-matched wild-type treated 88
Ablation of p16-positive cells 12 Kidney, heart and adipocyte metrics and function 12 months of age and age-matched vehicle control 62
Ablation of p16-positive cells 24 Vasculature function Age-matched vehicle control 82
Ablation of p16-positive cells 12 Cartilage degeneration Age-matched vehicle control 67
Ablation of p16-positive cells 24 Fat accumulation in liver Age-matched vehicle control 64
Ablation of p16-positive cells 12, 20 Bone metrics and loss Age-matched vehicle control 69
Ablation of p16-positive cells >25 Renal function Age-matched vehicle control 61
Dasatinib + quercetin 24 Cardiac metrics and function Age-matched vehicle control 68
Dasatinib + quercetin 24 Vasculature function Age-matched vehicle control 82
Dasatinib + quercetin 24 Fat accumulation in the liver Age-matched vehicle control 64
Dasatinib + quercetin 20 Bone metrics and loss Age-matched vehicle control 69
Dasatinib + quercetin 20 Physical performance Age-matched vehicle control 70
ABT263 21–22 Immune system (HSC) and muscle (MuSC) function 2 months of age and age-matched vehicle control 65
UBX0101 2–3, 19–20 Cartilage regeneration Age-matched vehicle control 67
FOXO4-DRI 24 Renal function and frailty Age-matched vehicle control 61
Cellular reprogramming
Transient reprogramming 12 Pancreas regeneration Age-matched vehicle control 105
Transient reprogramming 12 Muscle regeneration Age-matched vehicle control 105

An extended version of this table with more details regarding the method of administration or procedure, duration of intervention, mouse strain and sex is available as Supplementary Table 1 HSC, haematopoietic stem cell; MuSC, muscle stem cell.

How does young blood revitalize aged organs and tissues? Young blood may contain pro-rejuvenation factors, or it could dilute or inhibit pro-ageing factors in aged blood (Fig. 2). The pro-rejuvenation effect of young blood on the liver, muscle and brain is less pronounced than the pro-ageing effect of aged blood on these tissues27, suggesting the presence of potent pro-ageing factors in aged blood. Indeed, systemic pro-ageing factors have been identified through heterochronic parabiosis, including eotaxin (also known as CCL11)21 and β2-microglobulin23. The levels of eotaxin and β2-microglobulin increase with age, and these factors inhibit neurogenesis and cognition in young mice21,23. Whether blocking pro-ageing blood factors improves tissue function in aged mice remains to be shown, but aged β2-microglobulin knockout mice exhibit enhanced neurogenesis and cognitive functions compared to age-matched wild-type mice21,23. Other systemic signalling pathways have been implicated in mediating the pro-ageing effect of aged blood16,17,28-31. For example, heterochronic parabiosis reverses the excessive Wnt signalling underlying the differentiation bias of aged muscle stem cells16,29. Furthermore, systemic attenuation of transforming growth factor-β signalling improves age-dependent decline in neurogenesis and myogenesis30, and inhibition of interferon signalling partially ameliorates neurogenesis and cognitive function in aged mice31. Thus, several pro-ageing factors have been identified in aged blood.

Fig. 2 ∣. Potential common mechanisms and target cells of the rejuvenation strategies.

Fig. 2 ∣

A comparison of the proposed underlying mechanisms of action and target cell types influenced by each of the rejuvenation strategies. These include subcellular mechanisms (for example, chromatin changes, induction of autophagy pathways and alteration in mitochondrial function), cellular functions (such as revival of stem cell populations, attenuation of the deleterious effects of senescent cells and changes in connective tissue cells (for example, endothelial cells, fibroblasts and adipocytes)) and intercellular features (for example, decrease in inflammation, perturbation of nutrient-sensing pathways and changes in blood factors). The circles below each feature are colour-coded for each rejuvenation strategy and represent the current level of evidence for the effect of the corresponding strategy on the feature. Solid/dark circles, strong evidence. Dotted/light circles, mostly indirect evidence. Question marks, no evidence as of now.

Identifying rejuvenation factors in young blood has been more difficult. Heterochronic parabiosis can restore the decreased Notch signalling that underlies the decline in muscle stem cell activation and number17,28, although the specific systemic factor (or factors) remains unclear. Growth/differentiation factor 11 (GDF11) was initially identified as a circulating factor whose levels decrease with age but are restored through heterochronic parabiosis18. Exogenous GDF11 can rejuvenate heart function18,32 and improve muscle and neural stem cell functions in aged mice20,22 (Table 1 and Supplementary Table 1). However, subsequent studies reported no beneficial effects of GDF11 on heart and muscle stem cell function33,34, and injection of GDF11 can induce cachexia35. Thus, although GDF11 may have beneficial effects under specific conditions, it is unlikely to be a universal mediator of rejuvenation. Another potential rejuvenating blood factor whose levels decrease with age is the hormone oxytocin36. Its systemic administration improves muscle regeneration by enhancing muscle stem cell activation and/or proliferation in aged mice36 (Table 1 and Supplementary Table 1). As oxytocin is known for its role in social bonding37, it could potentially link social environment and ageing. Finally, TIMP2, a metalloproteinase inhibitor, was identified in human umbilical cord plasma26 and its levels have been shown to decrease with age in both mice and humans26. Injection of human cord blood induces hippocampal neurogenesis and improves learning and memory in naturally aged mice26, effects that are attenuated by TIMP2 depletion26. Moreover, administration of exogenous TIMP2 can improve cognitive function in aged mice, pointing to TIMP2 as a key rejuvenating factor26 (Table 1 and Supplementary Table 1).

These studies suggest the presence of both pro-ageing and anti-ageing factors in the blood, which can be targeted to reverse age-related decline in multiple tissues. However, many open questions remain. Which cell types secrete these factors and could these cells be targeted to achieve similar effects? Do circulating factors drive rejuvenation of all tissues or do they have tissue-specific action? Comprehensive analysis of the response of multiple organs to blood factors will be required to address this question. Testing the interaction between individual factors, including pro-ageing and rejuvenation factors, could identify the main contributors and allow for combinatorial treatments to revitalize tissues and organs. A central question is whether young blood or specific blood factors can extend organismal lifespan. Although initial studies using young blood in aged mice or GDF11 in progeroid mice reported little effect on overall lifespan38,39, a thorough investigation of the effect of blood factors on mammalian lifespan will be important.

Metabolic-induced rejuvenation

Long-term dietary restriction extends healthspan and lifespan across several species12,40. Less-restrictive diet regimens and drugs that mimic the metabolic effects of dietary restriction also have beneficial effects on lifespan5,8-10,41,42. Until recently, it was unclear whether these interventions could reverse ageing features in aged individuals. Initial studies on short-term dietary restriction (5 days to 12 weeks) in middle-aged or old-aged mice revealed improved function in multiple tissues, including muscle, bone, liver, brain, vasculature and immune system5,7,43 (Fig. 1, Table 1 and Supplementary Table 1), consistent with the possibility that dietary interventions could indeed reverse functional decline. Here, we focus on dietary interventions or mimics that are initiated at middle age or later and on their potential rejuvenation effects on ageing hallmarks.

The periodic fasting-mimicking diet (FMD) consists of cycles of very low caloric intake for 4 days, repeated twice per month, with ad libitum feeding in between5. When initiated in 16-month-old mice, FMD reverses age-associated haematopoietic differentiation bias, increases hippocampal neurogenesis and improves hippocampus-dependent memory5 (Table 1 and Supplementary Table 1). FMD also increases median lifespan and decreases cancer incidence and inflammatory diseases, including ulcerative dermatitis5. Some of the beneficial effects of FMD are probably mediated by an increased proliferative capacity and number of stem cells5. The refeeding portion of FMD may play a key role in this, as it results in a boost in cell proliferation5. Whether FMD also improves tissue function by selecting against dysfunctional cells is unclear. The proliferation boost following refeeding may favour youthful cells, diluting out damaged ones and improving overall tissue function. Given that FMD (and the ketogenic diet, discussed below) reduces cancer incidence5,9, such regimens may also select against cancerous or precancerous cells.

The ketogenic diet involves the same caloric intake as a normal diet but with reduced carbohydrate consumption. This diet mimics many of the metabolic changes occurring in mice under dietary restriction or fasting8,9,44. Both fasting and a ketogenic diet decrease blood glucose levels and increase ketone body levels and fatty acid oxidation8,9,44 (Table 1 and Supplementary Table 1). Interestingly, alternating between a ketogenic and a control diet weekly in middle-aged mice improves recognition memory and midlife survival8. A non-cyclic ketogenic diet also increases median lifespan and improves motor function in aged mice while decreasing cancer incidence9. Although not explicitly stated in these studies, some measurements are similar or better post-treatment compared to those at the time of treatment initiation, suggesting not only a delay but also a reversal of the measured features9,10. Thus, manipulating diet content may constitute an effective approach for reversing ageing hallmarks and may be easier to implement in humans than long-term dietary restriction.

How do these diet regimens rejuvenate aged tissues? Nutrient-sensing pathways, including mTOR and insulin–IGF signalling, could play a key role3,45-47 (Fig. 2). Periodic FMD was proposed to act by reducing insulin–IGF1 signalling and inhibiting the activity of mTOR and protein kinase A6. Short-term treatment (6 weeks) with rapamycin, an mTOR inhibitor, improves haematopoietic stem cell function in aged mice (although not to the level of 2-month-old mice) and extends lifespan48 (Table 1 and Supplementary Table 1). The insulin and mTOR signalling pathways are also known to regulate autophagy3,45-47, and FMD can indeed counteract the decline in autophagy-related proteins in ageing muscle5. This points to an important role of mTOR in mediating the beneficial effects of these regimens and raises the possibility that mTOR inhibitors could be used to rejuvenate ageing tissues. Although the effect of rapamycin on lifespan is well established45, whether it is a rejuvenating compound remains debated. A comprehensive assessment of ageing phenotypes following long-term (1 year) treatment of young and aged mice showed that rapamycin improves several features, including memory and learning49. However, it also ameliorates some of these features in young mice, suggesting that it may have age-independent positive effects49.

Similarly to periodic FMD, the ketogenic diet also inhibits mTOR and insulin–IGF signalling8,9. Interestingly, although a short-term ketogenic diet (1 month) does affect the expression of genes related to insulin signalling and fatty acid synthesis, an extended ketogenic diet (14 months of cyclic diet) does not affect these genes8. Thus, repeated cycles may become less effective on signalling pathways8. Ketogenic effects could be mediated by increased circulating β-hydroxybutyrate levels, a ketone that inhibits histone deacetylases and may thereby link metabolism, epigenetics and rejuvenation8,9. Hence, β-hydroxybutyrate could represent an effective longevity and rejuvenating compound50,51.

Other nutrient-sensing pathways could also be involved in the rejuvenation effects of dietary regimens (Table 1 and Supplementary Table 1). For example, metformin, which increases AMPK activity42,47, preserves mitochondrial function and decreases inflammation when administered starting at middle age41. Resveratrol, which can activate sirtuins (and other nutrient-responsive pathways), also improves cognitive and renal function and reduces inflammation in rodents when initiated at mid-to-late life42,52-54. Whether these improvements represent a true reversal of pre-existing ageing phenotypes remains an open question.

Ablation of senescent cells to restore tissue youthfulness

Cellular senescence is a cell-intrinsic mechanism induced by stress that prevents propagation of damaged cells55,56. Initially identified as a barrier against tumour development56, senescence is now known to be involved in tissue remodelling during embryogenesis57,58, wound healing59,60 and ageing61-70. Senescence markers include senescence-associated β-galactosidase activity, the cell-cycle inhibitors p16INK4a and p21CIP1, and many secreted inflammatory factors (collectively referred to as the senescence-associated secretory phenotype (SASP))55,56,71. Senescent cells are heterogeneous72,73 and do not always exhibit all markers, and, conversely, some markers are also present in non-senescent cells71,74. Senescent cells accumulate in ageing tissues across organisms, including primates and rodents, and in age-related pathologies, such as atherosclerosis and Alzheimer’s disease75-79. Accordingly, senescence has long been thought to contribute to organismal ageing56, although whether it is a cause or consequence is only starting to be resolved. Indeed, mouse models and compounds that trigger senescent cell elimination have revealed that targeting senescent cells can reverse or delay aspects of the ageing process61-70 (Fig. 1).

The first evidence that senescent cells can actively contribute to ageing came from genetically modified mice that allow for inducible elimination of p16-positive cells in the context of a progeroid disease63. In INK-ATTAC transgenic mice that express a drug-inducible form of caspase 8 under the Cdkn2a (which encodes P16Ink4a) promoter, drug administration triggers caspase-8-mediated apoptosis in p16-positive cells63. In a progeroid mouse model (BubR1), caspase-8-mediated ablation of p16-positive cells starting from early life delays the onset of age-associated features, including loss of fat and skeletal muscle and cataract development63. Even later in life, ablation of p16-positive cells reduces age-associated fat and skeletal muscle loss63. Follow-up studies in naturally ageing mice showed that the removal of p16-positive cells starting from 12 months of age (midlife) through to 18 months of age attenuates age-associated decline in adipocyte, kidney and heart function62 (Table 1 and Supplementary Table 1). Importantly, the removal of p16-positive cells from midlife to end of life extends median lifespan by 24–27%62. Similar health benefits were observed with another model. In p16-3MR mice that express thymidine kinase from herpes simplex virus under the Cdkn2a promoter, administration of the thymidine kinase substrate ganciclovir initiates apoptosis of p16-positive cells59. In both mouse models, apoptosis induction in p16-positive cells late in life for at least 3 weeks improves liver, kidney, bone and adipocyte metrics and function61,64,69. Although not specifically stated, some metrics seem to be better after treatment than at initiation61,64, suggesting that senescent cell ablation may reverse ageing features. However, high p16 expression has also been observed in non-senescent cells, notably macrophages71,74. Hence, the beneficial effects of this intervention might partially be due to targeting macrophages, which are known to change with age1,2,80. Despite the promise of these initial findings, more remains to be learned about the optimal times for treatment initiation and duration for maximal effects, and about specificity to senescent versus immune cells.

These initial proof-of-concept studies spurred the field to identify compounds that can relatively specifically kill senescent cells based on their unique molecular profiles. Several classes of such ‘senolytic’ drugs have been identified, including Bcl protein family inhibitors (for example, navitoclax, also known as ABT263)65, kinase inhibitors (for example, dasatinib and quercetin)68, heat shock protein 90 inhibitors (for example, 17-DMAG)66 and inhibitors of the p53–MDM2 interaction (for example, UBX0101)67,81. Dasatinib and quercetin68 and 17-DMAG66 improve healthspan in the Ercc1−/− progeroid mouse model. In naturally aged mice, senolytics enhance cardiovascular, vascular, bone, liver and physical functions (dasatinib and quercetin)64,68,69,82, revitalize haematopoietic and muscle stem cell populations (ABT263)65, enhance cartilage regeneration (UBX0101)67 and even extend median lifespan (dasatinib and quercetin)70 (Table 1 and Supplementary Table 1). A forkhead box protein O4 peptide (FOXO4-DRI) also has senolytic effects. This peptide blocks the sequestration of p53 by FOXO4, which seems to be senescence specific, thus allowing p53 activation and cell death in senescent cells61. FOXO4-DRI restores fitness, fur density and kidney function in both progeroid (XpdTTD/TTD) and naturally aged mice61 (Table 1 and Supplementary Table 1). Whether FOXO4-DRI acts on all types of senescent cells without targeting healthy cells, a common challenge for senolytic drugs14,68,83-85, remains to be determined. Many senolytics were initially identified as cancer drugs because cancer cells exploit similar anti-apoptotic pathways, notably overexpression of Bcl family proteins84,86. Thus, some beneficial effects of senolytics may originate from the elimination of precancerous and cancerous cells.

How does the removal of senescent cells rejuvenate tissues and extend lifespan? Senescence could contribute to the decline in tissue homeostasis and function by inducing a permanent cell-cycle arrest in proliferative cell populations. Senescence of reparative stem and progenitor cells may lead to a decline in tissue regenerative potential. Senescence could also act through SASP, which promotes local and systemic inflammation55,56. SASP factors could contribute to stem cell exhaustion or dysfunction, infiltration and alteration of immune cells, insulin resistance, damage of tissue structure and even propagation of the senescent phenotype in neighbouring cells56,87. Elimination of senescent cells can revive stem cell populations in naturally aged mice65,88 (Fig. 2), and p16 depletion resets ageing features in aged muscle stem cells89. Moreover, SASP inhibition by the Janus kinase 1/2 inhibitor ruxolitinib reduces systemic and adipose tissue inflammation and increases insulin sensitivity in naturally aged mice69,88,90. Senescent cell removal can delay cancer development, which could be a source of the observed lifespan extension in mice62. Elucidating the mechanisms by which senolytics ameliorate tissue function will be important in identifying additional senolytic compounds and in determining how best to use them. Importantly, senescent cells can have beneficial effects, for example, by facilitating tissue repair after injury and preventing tissue fibrosis59,60,91. Identifying mechanisms that distinguish between the beneficial and harmful effects of senescence could help to identify therapeutic strategies to specifically target the latter.

Reprogramming back to a youthful state

Cellular reprogramming is the conversion of terminally differentiated somatic cells into induced pluripotent stem cells (iPSCs)92, for instance by the expression of the transcription factors OCT4 (also known as POU5F1), SOX2, KLF4 and MYC (OSKM)92. Cellular reprogramming allows for the generation of in vitro models to study ageing and age-associated diseases and the development of autologous stem cell therapies to replace ageing tissues93-95. Reprogramming also resembles to some extent the process of fertilization, during which the chronological age of the parent cells is effectively reset such that the resulting offspring has a normal lifespan96. Hence, cellular reprogramming has emerged as a potential rejuvenation strategy15,96.

Reprogramming to pluripotency can erase several ageing features in vitro. iPSCs derived from aged cells show extended telomeres, improved mitochondrial morphology, number and fitness (ATP production and membrane potential) and restored nuclear morphology15,95,97,98. iPSC reprogramming of aged cells also resets heterochromatin marks and transcriptomic profiles15,95,97,99. After re-differentiation of these iPSCs into neurons or fibroblasts, transcriptomic changes, improvements in nucleocytoplasmic compartmentalization, nuclear morphology and (in the case of fibroblasts) proliferative potential largely remain in the rejuvenated state95,97,99. This suggests that the youthful state is not exclusive to pluripotency and can persist after re-differentiation. Although most age-associated phenotypes tested are reversed by in vitro reprogramming, iPSCs generated from aged human cells can retain a DNA methylation signature of their age, which can be erased with additional passaging100. Thus, some features of ageing may be harder to rejuvenate than others, and some aspects, such as pre-existing genetic mutations, cannot be reverted94,100. The ability to rejuvenate ageing traits may be specific to reprogramming to a pluripotent state because direct reprogramming to a differentiated state (for example, neurons) was less effective at erasing ageing marks99. Future studies should explore the extent and time course of molecular rejuvenation by iPSC reprogramming, to determine whether there is dependency between different age-associated features.

Recent studies using mouse models of doxycycline-inducible reprogramming factor (OSKM) expression have demonstrated that somatic cells can be reprogrammed to pluripotency in vivo101-104, suggesting that the rejuvenating effects of cellular reprogramming might be recapitulated in an organism. A major limitation of initial studies was that persistent expression of OSKM led to teratoma formation101-103. Thus, an important step was to determine whether the rejuvenating aspect of reprogramming could be uncoupled from its dedifferentiating, teratoma-inducing properties96. Interestingly, this uncoupling was recently shown to be possible105. Short-term OSKM induction (‘partial reprogramming’) in fibroblasts from progeroid mice (LmnaG608G) erased features of ageing, including DNA damage, dysregulation of histone marks, expression of senescence-associated genes and nuclear envelope abnormalities105. When applied in vivo, cyclic partial reprogramming (2-day induction with 5-day withdrawal) starting at 8 weeks of age extended both healthspan and lifespan (median ~30%, maximum ~20%) of these mice, without teratoma or cancer development105 (Fig. 1). In vivo partial reprogramming applied to naturally ageing mice at midlife also improved glucose tolerance and the regenerative capacity of muscle and the pancreas after injury105 (Table 1 and Supplementary Table 1). These observations underscore the potential of cellular reprogramming to rejuvenate cells and tissues in vivo, although more work is needed in the context of naturally aged mice. Indeed, some of the positive OSKM effects in muscle at midlife could be age independent, as the regenerative potential of muscle has not declined yet at this stage of life36,106. Whether partial reprogramming can reverse tissue decline in the absence of injury or disease and/or extend lifespan in naturally aged mice also remains to be determined.

How does cellular reprogramming rejuvenate aged cells and tissues? At the molecular level, epigenetic remodelling is a key factor in iPSC reprogramming107,108, and histone modifications have been proposed to mediate the rejuvenating effects of partial reprogramming105 (Fig. 2). At the tissue level, partially reprogrammed mice have increased numbers of muscle stem cells after injury105. Hence, enhanced regenerative capacity and stem cell function could contribute to the lifespan extension observed in the context of premature ageing105. Reprogramming could also act by eliminating dysfunctional cells in tissues or by diluting them through proliferation of healthy cells. The extent to which rejuvenating effects persist after in vivo reprogramming remains an important direction for future studies. Although some reprogramming-induced epigenetic and transcriptomic remodelling persists following doxycycline withdrawal101, the increase of histone 3 lysine 9 trimethylation (H3K9me3) levels reverts within 8 days of withdrawal in vitro105. Thus, whether transient reprogramming leads to transient or persistent rejuvenation remains to be determined.

Common or distinct mechanisms of rejuvenation

One key question is whether the four rejuvenation strategies described above share modes of action or whether they use distinct mechanisms (Fig. 2). Common pathways could be harnessed to induce rejuvenation more directly, whereas differing ones could be targeted in combination to enhance it.

Inflammation.

Inflammation could be directly or indirectly affected by most rejuvenation strategies. Heterochronic parabiosis reduces inflammatory factors and pathways, such as eotaxin and interferon signalling21,31. FMD and dietary restriction (DR)-mimicking drugs have anti-inflammatory effects by suppressing the onset of senescence and the secretion of pro-inflammatory cytokines109-111. Senolytics could exert their beneficial effects by reducing inflammation, as senescent cells contribute to inflammation through SASP56,87. Finally, although age-associated activation of nuclear factor-κB signalling impairs cellular reprogramming112, activation of innate immunity and inflammatory factors, such as interleukin-6 (IL-6), promote reprogramming102,104,113-115. These observations highlight inflammation as a critical target for rejuvenation strategies. Chronic inflammation (‘inflammaging’) has emerged as a key feature of ageing and age-associated diseases1,2,116, and its genetic and pharmacological targeting has been shown to extend healthspan and lifespan across multiple species117-121. Interestingly, stimulation or blocking of hypothalamic nuclear factor-κB activity was shown to accelerate or decelerate ageing, respectively122, suggesting a potential key role of the hypothalamus in modulating inflammation and ageing. Future studies should aim at investigating the interplay between rejuvenation strategies and inflammation, and exploring potential synergistic effects of rejuvenating compounds with anti-inflammatory drugs.

Nutrient-sensing pathways.

The insulin–IGF1, mTOR and AMPK pathways have been extensively studied in the context of longevity1-3,45-47 and are key candidates for relaying rejuvenating effects. The anti-ageing diets discussed inhibit mTOR and/or elicit a drop in circulating insulin and IGF1 levels5,8,9. DR-mimicking drugs also inhibit insulin–IGF1 and mTOR signalling and activate AMPK41,49,123. However, evidence for the involvement of these pathways in heterochronic parabiosis, the elimination of senescent cells and cellular reprogramming is mostly circumstantial. The shared circulatory system and organs in parabiosis may affect glucose–insulin homeostasis and IGF1 signalling124. Moreover, the IGF1 and mTOR pathways promote senescent cell survival and regulate SASP125-127, whereas AMPK pathway activation suppresses the development of senescence128. Finally, insulin–IGF1 signalling inhibits reprogramming15,129,130, although the role of AMPK in cellular reprogramming is still debated15. An intriguing possibility is that nutrient-sensing pathways may be more important for delaying ageing than reversing it.

Epigenomic remodelling.

The epigenomic landscape of a cell reflects not only its identity but also its health and biological age131-133. Senescent cells exhibit a characteristic chromatin state134,135, and their secreted factors (for example, IL-6) have been shown to induce epigenomic changes136,137. The rejuvenating effect of cellular reprogramming has been proposed to occur through epigenomic remodelling105. Moreover, dietary interventions and DR-mimicking drugs affect the epigenome131,138,139, although whether these changes are necessary for rejuvenating effects is unclear. Finally, while chromatin changes have not yet been reported in the context of heterochronic parabiosis, chromatin changes could relay some effects13. Whether restoring a youthful epigenome holds the key to a prolonged rejuvenated state is a compelling question.

Autophagy.

Autophagy, which includes the process of delivering damaged proteins and organelles to lysosomes for degradation, is key for cellular homeostasis140 and could play an important role in mediating rejuvenation. Most diet regimens and DR-mimicking drugs induce autophagy5,140,141, and the blood factor GDF11 was shown to enhance this process20. Senescent cell ablation could eliminate autophagy-deficient cells142,143. Finally, autophagy is also induced early in the reprogramming process129. Whether autophagy is necessary for the rejuvenation effects of cellular reprogramming remains unclear15,129, but reactivation of the lysosome–autophagy pathway in aged stem cells improves their function144-146. These observations suggest a link between the lysosome–autophagy pathway and rejuvenation strategies, but the extent to which autophagy promotes rejuvenation remains to be explored.

Mitochondria.

Mitochondrial function could also be central to rejuvenation strategies. Cellular reprogramming increases mitochondrial fitness98,147 and GDF11 can improve mitochondrial morphology and function20. Senescent cells have dysfunctional mitochondria with increased generation of reactive oxygen species, which in turn promote SASP and can induce senescence in neighbouring cells148-150. Mitochondria in senescent cells were recently suggested to have reduced ability to metabolize fatty acids, contributing to increased hepatic fat deposition with age and a decline in liver function64. Hence, the removal of senescent cells with poor mitochondrial function could be beneficial by reducing reactive oxygen species levels in the microenvironment and perhaps also by improving overall mitochondria function in ageing tissues and organs. However, rejuvenation strategies could also act by reducing mitochondrial function. Indeed, reduced mitochondrial activity extends lifespan in Caenorhabditis elegans, Drosophila and mice151-155. In addition, metformin, which inhibits mitochondrial function141, can extend healthspan and/or lifespan in multiple organisms41,141. Future studies should explore how this organelle relays the rejuvenating effects of these different strategies.

These observations suggest that the four rejuvenation strategies could act through common molecular pathways. However, the degree to which these pathways are modulated and whether each strategy targets them directly or indirectly remain unclear. Investigating the regulation and sequential order of these pathways following each intervention will help to identify the mechanisms that are critical for restoring youthfulness and that could be targeted for greater effect. Different rejuvenation approaches could also act via diverse mechanisms, which could be combined to achieve synergistic effects. Broader conceptual questions also remain: is the rejuvenation process the direct opposite sequence of events that lead to ageing? Do rejuvenation strategies target the root cause of ageing or simply its consequences? Can these interventions affect overall lifespan?

Target cells for rejuvenation

Which cell types are primarily targeted by rejuvenation strategies and mediate their beneficial effects? Adult stem cells are an attractive candidate as they provide a renewable source of cells to repair damaged tissues (Fig. 2). Indeed, most rejuvenation approaches also improve stem cell functions5,17,21,22,25,65,105,110,156, although whether these effects are direct or indirect remains unclear. The inherent plasticity of stem cells may make them more susceptible to the rejuvenating effects of cellular reprogramming, for example.

Stem cell state may also dictate susceptibility to ageing and rejuvenation. For example, quiescent stem cells exhibit increased age-related features compared to actively proliferating stem cells146,157, raising the possibility that quiescent cells might benefit more from rejuvenation strategies. In fact, a proliferative state could itself reset ageing features (for example, DNA damage and protein aggregates) in stem cell populations146,157. In addition, these rejuvenation strategies could indirectly affect stem cells. For example, young blood was proposed to enhance neurogenesis in aged mice by improving endothelial cells and thereby the vasculature of the neural stem cell niche22. Moreover, although the senolyte ABT293 is thought to improve aged haematopoietic and muscle stem cells by eliminating senescent stem cells65, it could also act by clearing senescent niche cells, such as endothelial cells and fibroblasts. In line with this notion, niche endothelial cells were shown to contribute to haematopoietic stem cell ageing, and transplantation of young endothelial cells could partially reverse these changes158. Thus, the primary target of rejuvenation approaches may be vascular and connective tissue cells. As these cells are present throughout the organism, targeting them may have broader organismal effects. Teasing apart the effects of rejuvenation strategies on different cell types and states will help efforts to improve tissue function and health and could identify strategies to simultaneously target both differentiated and stem cell populations for enhanced treatments.

Other attractive candidate cells for rejuvenation are senescent cells. Beyond their direct elimination by genetic means or senolytics, senescent cells may also be targeted by other rejuvenation strategies. The pro-ageing factor eotaxin21 has been associated with senescence159, potentially linking the beneficial effect of senescent cell ablation to changes in systemic factors. Moreover, FMD, DR and DR-mimicking drugs suppress senescence onset and pro-inflammatory cytokine levels109-111,160. Although speculative, it is also plausible that the proliferation bursts induced by FMD or partial reprogramming could dilute and/or trigger senescent cell clearance. Indeed, many of the age-associated features that are reverted by partial reprogramming are related to senescence105. Cellular reprogramming has been suggested to rejuvenate senescent cells97. However, the relationship between reprogramming and senescence is complex. Reprogramming factors can trigger cellular senescence161,162; conversely, senescence promotes cellular plasticity of neighbouring cells through SASP (for example, IL-6)102,104,115,163. In line with these observations, induction of reprogramming factors for 7 days results in more teratomas in aged mice than in young mice102-104, possibly due to the presence of senescent cells in aged tissues. Finally, senescent cell removal using senolytic drugs or an inducible genetic system decreases in vivo reprogramming efficiency115. It will be interesting to elucidate the interplay between senolytic and reprogramming strategies for rejuvenation.

Potential trade-offs of rejuvenation

Ageing disrupts the balance of key biological processes that maintain organismal homeostasis and function. Hence, reversing it is not as simple as turning off these processes, but rather involves the need to restore a balance. For example, although age-associated senescence and/or chronic inflammation could impair tissue function, they are also critical for normal tissue repair and remodelling59,60. Accordingly, counteracting senescence and/or inflammation could reduce the ability of the organism to perform these processes (Fig. 1). Indeed, elimination of senescent cells impedes tissue repair and promotes tissue-specific fibrosis59,60,91. Similarly, DR-related interventions impair the immune response to infections and reduce wound healing164, although refeeding after DR or DR-mimicking drugs can restore or even potentiate these responses164,165. DR regimens, when started too early, can also interfere with growth and fecundity and lead to amenorrhea and osteoporosis12. Importantly, excessive perturbation of a specific feature may ultimately lead to tumorigenesis and cancer progression. As senescence is a critical barrier against tumorigenesis56, preventing its induction could increase cancer risk. Similarly, sustained expression of reprogramming factors could lead to tumour formation101. Senescent cells also exploit anti-apoptotic pathways, such as Bcl-2, that are important for the survival of healthy cells (for example, lymphocytes and platelets)166-168. Consequently, compounds that are used to target senescent cells (for example, pan-Bcl inhibitors) are also associated with gastrointestinal symptoms and haematopoietic system toxicity83,84. Hence, the risk/benefit ratio of these rejuvenation strategies must be taken into account before considering them as a viable anti-ageing treatment.

Future perspectives

There is now compelling evidence that the ageing process is plastic and that it is possible to revive aged cells and tissues. Although the four strategies discussed here have received much attention in recent years, other approaches may also turn out to have rejuvenating effects. Genetic perturbations such as the expression of telomerase in middle-aged and old-aged mice improves healthspan (for example, insulin sensitivity and osteoporosis) and extends median lifespan169. Similarly, life-long increased dosage of p16INK4 and p53 can have beneficial effects to counter ageing170-172. Hence, inducible telomerase, p16 or p53 expression later in life could be future rejuvenation strategies. Environmental interventions that have benefits on healthspan and lifespan could also be leveraged for rejuvenation. For example, exercise improves hippocampal neurogenesis and muscle function in aged rodents173-175. Lowering core body temperature extends lifespan in invertebrates and African killifish3,176-178 and even in mice179. Finally, the transfer of young microbiome in middle-aged killifish was recently shown to extend both healthspan and lifespan180. However, whether these potential strategies revert ageing hallmarks or delay the appearance of such characteristics remains to be tested. It will also be interesting to determine whether key organs or systems, such as the hypothalamus, orchestrate ageing in a centralized manner by integrating environmental inputs and secreting systemic factors36,122,181. These systems could then be targeted to achieve whole-organism rejuvenation.

The question also emerges of whether rejuvenation interventions, which were mainly tested in mice, may benefit human health and longevity (Fig. 1). Metabolic approaches have reached furthest in testing this possibility and have shown promise in benefiting humans. FMD in individuals ranging from 20 to 70 years of age was shown to improve physiological readouts that are altered with age, including body weight, blood pressure, cholesterol and IGF1 levels5,182. FMD and DR-mimicking drugs, such as metformin and rapamycin, can improve risk factors associated with age-related diseases, such as cancer, diabetes and cardiovascular disease182-185. Clinical trials are underway using metformin and rapamycin to target ageing141 (ClinicalTrials.gov identifiers: NCT02432287 and NCT02874924) and rapamycin analogues are being tested in the elderly in the context of response to vaccination165 and respiratory tract infection (ClinicalTrials.gov identifier: NCT03373903). Currently, there are no data showing beneficial effects of blood factors, senolytic drugs or reprogramming in humans. However, the levels of the pro-ageing blood factors eotaxin and β2-microglobulin are increased in the plasma of elderly humans21,23 and the rejuvenation factor TIMP2 is enriched in human umbilical plasma26. Moreover, most senolytic drugs identified can eliminate human senescent cells in vitro61,65,66,68,85,186. Similarly, cellular reprogramming can revert ageing features of human cells in vitro95,97-100,105, raising the possibility that these approaches may also prove beneficial for human ageing. Indeed, some of these approaches are now being explored in the context of human age-associated diseases. For instance, young blood is being tested in Alzheimer’s disease (ClinicalTrials.gov identifier: NCT02256306). Although the initial trial showed only a minor improvement187, larger trials are underway to better assess efficacy. Several senolytics are currently used in the clinic as anticancer drugs84,86 and are being tested on chronic kidney disease (ClinicalTrials.gov identifier: NCT02848131) and osteoarthritis (ClinicalTrials.gov identifier: NCT03513016). Initial findings are encouraging, but many challenges remain before these strategies can be used successfully in the clinic. The optimization of therapeutic dosage with minimal side effects will be key to translational efforts. It will also be critical to establish reasonable end points and robust biomarkers of healthy ageing to assess intervention efficacy.

These studies provide compelling evidence that the ageing process is malleable and that it is possible to revive aged cells, tissues and organs. They also raise the exciting possibility of translation to address human ageing and age-associated diseases. The coming years will undoubtedly see exciting developments in ongoing efforts to better understand, delay and potentially reverse ageing.

Supplementary Material

Table 1

Acknowledgements

We apologize to those colleagues whose work we could not cite owing to space limitations. We thank C. Kenyon, P. Singh, J. Vos, M. Quarta and A. Colville for helpful feedback on the manuscript. This work was supported by the Stanford Graduate Fellowship (L.X.) and a generous philanthropic gift from M. Barakett and T. Barakett.

Footnotes

Competing interests

The authors declare no competing interests.

Supplementary information is available for this paper at https://doi.org/10.1038/s41556-018-0206-0.

Online content

Any methods, additional references, Nature Research reporting summaries, source data, statements of data availability and associated accession codes are available at https://doi.org/10.1038/s41556-018-0206-0.

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Lopez-Otin C, Blasco MA, Partridge L, Serrano M & Kroemer G The hallmarks of aging. Cell 153, 1194–1217 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Kennedy BK et al. Geroscience: linking aging to chronic disease. Cell 159, 709–713 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Kenyon CJ The genetics of ageing. Nature 464, 504–512 (2010). [DOI] [PubMed] [Google Scholar]
  • 4.Gems D & Partridge L Genetics of longevity in model organisms: debates and paradigm shifts. Annu. Rev. Physiol 75, 621–644 (2013). [DOI] [PubMed] [Google Scholar]
  • 5.Brandhorst S et al. A periodic diet that mimics fasting promotes multi-system regeneration, enhanced cognitive performance, and healthspan. Cell Metab. 22, 86–99 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Cheng CW et al. Fasting-mimicking diet promotes Ngn3-driven β-cell regeneration to reverse diabetes. Cell 168, 775–788.e12 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Cerletti M, Jang YC, Finley LW, Haigis MC & Wagers AJ Short-term calorie restriction enhances skeletal muscle stem cell function. Cell Stem Cell 10, 515–519 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Newman JC et al. Ketogenic diet reduces midlife mortality and improves memory in aging mice. Cell Metab. 26, 547–557 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Roberts MN et al. A ketogenic diet extends longevity and healthspan in adult mice. Cell Metab. 26, 539–546 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Harrison DE et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392–395 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Longo VD et al. Interventions to slow aging in humans: are we ready? Aging Cell 14, 497–510 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.de Cabo R, Carmona-Gutierrez D, Bernier M, Hall MN & Madeo F The search for antiaging interventions: from elixirs to fasting regimens. Cell 157, 1515–1526 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Conboy MJ, Conboy IM & Rando TA Heterochronic parabiosis: historical perspective and methodological considerations for studies of aging and longevity. Aging Cell 12, 525–530 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.de Keizer PL The fountain of youth by targeting senescent cells? Trends Mol. Med 23, 6–17 (2017). [DOI] [PubMed] [Google Scholar]
  • 15.Mahmoudi S & Brunet A Aging and reprogramming: a two-way street. Curr. Opin. Cell Biol 24, 744–756 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Brack AS et al. Increased Wnt signaling during aging alters muscle stem cell fate and increases fibrosis. Science 317, 807–810 (2007). [DOI] [PubMed] [Google Scholar]
  • 17.Conboy IM et al. Rejuvenation of aged progenitor cells by exposure to a young systemic environment. Nature 433, 760–764 (2005). [DOI] [PubMed] [Google Scholar]
  • 18.Loffredo FS et al. Growth differentiation factor 11 is a circulating factor that reverses age-related cardiac hypertrophy. Cell 153, 828–839 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Baht GS et al. Exposure to a youthful circulation rejuvenates bone repair through modulation of β-catenin. Nat. Commun 6, 7131 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Sinha M et al. Restoring systemic GDF11 levels reverses age-related dysfunction in mouse skeletal muscle. Science 344, 649–652 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Villeda SA et al. The ageing systemic milieu negatively regulates neurogenesis and cognitive function. Nature 477, 90–94 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Katsimpardi L et al. Vascular and neurogenic rejuvenation of the aging mouse brain by young systemic factors. Science 344, 630–634 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Smith LK et al. β2-Microglobulin is a systemic pro-aging factor that impairs cognitive function and neurogenesis. Nat. Med 21, 932–937 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Ruckh JM et al. Rejuvenation of regeneration in the aging central nervous system. Cell Stem Cell 10, 96–103 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Villeda SA et al. Young blood reverses age-related impairments in cognitive function and synaptic plasticity in mice. Nat. Med 20, 659–663 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Castellano JM et al. Human umbilical cord plasma proteins revitalize hippocampal function in aged mice. Nature 544, 488–492 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Rebo J et al. A single heterochronic blood exchange reveals rapid inhibition of multiple tissues by old blood. Nat. Commun 7, 13363 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Conboy IM & Rando TA The regulation of Notch signaling controls satellite cell activation and cell fate determination in postnatal myogenesis. Dev. Cell 3, 397–409 (2002). [DOI] [PubMed] [Google Scholar]
  • 29.Brack AS, Conboy IM, Conboy MJ, Shen J & Rando TA A temporal switch from notch to Wnt signaling in muscle stem cells is necessary for normal adult myogenesis. Cell Stem Cell 2, 50–59 (2008). [DOI] [PubMed] [Google Scholar]
  • 30.Yousef H et al. Systemic attenuation of the TGF-p pathway by a single drug simultaneously rejuvenates hippocampal neurogenesis and myogenesis in the same old mammal. Oncotarget 6, 11959–11978 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Baruch K et al. Aging-induced type I interferon response at the choroid plexus negatively affects brain function. Science 346, 89–93 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Poggioli T et al. Circulating growth differentiation factor 11/8 levels decline with age. Circ. Res 118, 29–37 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Smith SC et al. GDF11 does not rescue aging-related pathological hypertrophy. Circ. Res 117, 926–932 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Egerman MA et al. GDF11 increases with age and inhibits skeletal muscle regeneration. Cell Metab. 22, 164–174 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Jones JE et al. Supraphysiologic administration of GDF11 induces cachexia in part by upregulating GDF15. Cell Rep. 22, 1522–1530 (2018). [DOI] [PubMed] [Google Scholar]
  • 36.Elabd C et al. Oxytocin is an age-specific circulating hormone that is necessary for muscle maintenance and regeneration. Nat. Commun 5, 4082 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lee HJ, Macbeth AH, Pagani JH & Young WS 3rd Oxytocin: the great facilitator of life. Prog. Neurobiol 88, 127–151 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Freitas-Rodriguez S, Rodriguez F & Folgueras AR GDF11 administration does not extend lifespan in a mouse model of premature aging. Oncotarget 7, 55951–55956 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Shytikov D, Balva O, Debonneuil E, Glukhovskiy P & Pishel I Aged mice repeatedly injected with plasma from young mice: a survival study. Biores. Open Access 3, 226–232 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Kapahi P, Kaeberlein M & Hansen M Dietary restriction and lifespan: lessons from invertebrate models. Ageing Res. Rev 39, 3–14 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Martin-Montalvo A et al. Metformin improves healthspan and lifespan in mice. Nat. Commun 4, 2192 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Bonkowski MS & Sinclair DA Slowing ageing by design: the rise of NAD+ and sirtuin-activating compounds. Nat. Rev. Mol. Cell Biol 17, 679–690 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Rippe C et al. Short-term calorie restriction reverses vascular endothelial dysfunction in old mice by increasing nitric oxide and reducing oxidative stress. Aging Cell 9, 304–312 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Meidenbauer JJ, Ta N & Seyfried TN Influence of a ketogenic diet, fish-oil, and calorie restriction on plasma metabolites and lipids in C57BL/6J mice. Nutr. Metab 11, 23 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Johnson SC, Rabinovitch PS & Kaeberlein M mTOR is a key modulator of ageing and age-related disease. Nature 493, 338–345 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zoncu R, Efeyan A & Sabatini DM mTOR: from growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol 12, 21–35 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Imai S & Guarente L NAD+ and sirtuins in aging and disease. Trends Cell Biol. 24, 464–471 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Chen C, Liu Y, Liu Y & Zheng P mTOR regulation and therapeutic rejuvenation of aging hematopoietic stem cells. Sci. Signal 2, ra75 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Neff F et al. Rapamycin extends murine lifespan but has limited effects on aging. J. Clin. Invest 123, 3272–3291 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Newman JC & Verdin E β-Hydroxybutyrate: a signaling metabolite. Annu. Rev. Nutr 37, 51–76 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Edwards C et al. d-β-hydroxybutyrate extends lifespan in C. elegans. Aging 6, 621–644 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Gocmez SS et al. Protective effects of resveratrol on aging-induced cognitive impairment in rats. Neurobiol. Learn. Mem 131, 131–136 (2016). [DOI] [PubMed] [Google Scholar]
  • 53.Kim EN et al. Resveratrol, an Nrf2 activator, ameliorates aging-related progressive renal injury. Aging 10, 83–99 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Pearson KJ et al. Resveratrol delays age-related deterioration and mimics transcriptional aspects of dietary restriction without extending life span. Cell Metab. 8, 157–168 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Hernandez-Segura A, Nehme J & Demaria M Hallmarks of cellular senescence. Trends Cell Biol. 28, 436–453 (2018). [DOI] [PubMed] [Google Scholar]
  • 56.Campisi J Aging, cellular senescence, and cancer. Annu. Rev. Physiol 75, 685–705 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Munoz-Espin D et al. Programmed cell senescence during mammalian embryonic development. Cell 155, 1104–1118 (2013). [DOI] [PubMed] [Google Scholar]
  • 58.Storer M et al. Senescence is a developmental mechanism that contributes to embryonic growth and patterning. Cell 155, 1119–1130 (2013). [DOI] [PubMed] [Google Scholar]
  • 59.Demaria M et al. An essential role for senescent cells in optimal wound healing through secretion of PDGF-AA. Dev. Cell 31, 722–733 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Krizhanovsky V et al. Senescence of activated stellate cells limits liver fibrosis. Cell 134, 657–667 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Baar MP et al. Targeted apoptosis of senescent cells restores tissue homeostasis in response to chemotoxicity and aging. Cell 169, 132–147 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Baker DJ et al. Naturally occurring p16Ink4a-positive cells shorten healthy lifespan. Nature 530, 184–189 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Baker DJ et al. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature 479, 232–236 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ogrodnik M et al. Cellular senescence drives age-dependent hepatic steatosis. Nat. Commun 8, 15691 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Chang J et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat. Med 22, 78–83 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Fuhrmann-Stroissnigg H et al. Identification of HSP90 inhibitors as a novel class of senolytics. Nat. Commun 8, 422 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Jeon OH et al. Local clearance of senescent cells attenuates the development of post-traumatic osteoarthritis and creates a pro-regenerative environment. Nat. Med 23, 775–781 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zhu Y et al. The Achilles’ heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell 14, 644–658 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Farr JN et al. Targeting cellular senescence prevents age-related bone loss in mice. Nat. Med 23, 1072–1079 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Xu M et al. Senolytics improve physical function and increase lifespan in old age. Nat. Med 24, 1246–1256 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sharpless NE & Sherr CJ Forging a signature of in vivo senescence. Nat. Rev. Cancer 15, 397–408 (2015). [DOI] [PubMed] [Google Scholar]
  • 72.Wiley CD et al. Analysis of individual cells identifies cell-to-cell variability following induction of cellular senescence. Aging Cell 16, 1043–1050 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Hernandez-Segura A et al. Unmasking transcriptional heterogeneity in senescent cells. Curr. Biol 27, 2652–2660.e4 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Hall BM et al. p16Ink4a and senescence-associated β-galactosidase can be induced in macrophages as part of a reversible response to physiological stimuli. Aging 9, 1867–1884 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Jeyapalan JC, Ferreira M, Sedivy JM & Herbig U Accumulation of senescent cells in mitotic tissue of aging primates. Mech. Ageing Dev 128, 36–44 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Liu Y et al. Expression of p16INK4a in peripheral blood T-cells is a biomarker of human aging. Aging Cell 8, 439–448 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Burd CE et al. Monitoring tumorigenesis and senescence in vivo with a p16INK4a-luciferase model. Cell 152, 340–351 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.McShea A, Harris PL, Webster KR, Wahl AF & Smith MA Abnormal expression of the cell cycle regulators p16 and CDK4 in Alzheimer’s disease. Am. J. Pathol 150, 1933–1939 (1997). [PMC free article] [PubMed] [Google Scholar]
  • 79.Childs BG et al. Senescent intimal foam cells are deleterious at all stages of atherosclerosis. Science 354, 472–477 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Linehan E & Fitzgerald DC Ageing and the immune system: focus on macrophages. Eur. J. Microbiol. Immunol 5, 14–24 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Villanueva MT Ageing: old bone removal. Nat. Rev. Drug Discov 16, 456 (2017). [DOI] [PubMed] [Google Scholar]
  • 82.Roos CM et al. Chronic senolytic treatment alleviates established vasomotor dysfunction in aged or atherosclerotic mice. Aging Cell 15, 973–977 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Schoenwaelder SM et al. Bcl-xL-inhibitory BH3 mimetics can induce a transient thrombocytopathy that undermines the hemostatic function of platelets. Blood 118, 1663–1674 (2011). [DOI] [PubMed] [Google Scholar]
  • 84.Wilson WH et al. Navitoclax, a targeted high-affinity inhibitor of BCL-2, in lymphoid malignancies: a phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. 11, 1149–1159 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Zhu Y et al. Identification of a novel senolytic agent, navitoclax, targeting the Bcl-2 family of anti-apoptotic factors. Aging Cell 15, 428–435 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Talpaz M et al. Dasatinib in imatinib-resistant Philadelphia chromosome-positive leukemias. N. Engl. J. Med 354, 2531–2541 (2006). [DOI] [PubMed] [Google Scholar]
  • 87.Nelson G et al. A senescent cell bystander effect: senescence-induced senescence. Aging Cell 11, 345–349 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Xu M et al. Targeting senescent cells enhances adipogenesis and metabolic function in old age. eLife 4, e12997 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Sousa-Victor P et al. Geriatric muscle stem cells switch reversible quiescence into senescence. Nature 506, 316–321 (2014). [DOI] [PubMed] [Google Scholar]
  • 90.Xu M et al. JAK inhibition alleviates the cellular senescence-associated secretory phenotype and frailty in old age. Proc. Natl Acad. Sci. USA 112, E6301–E6310 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Schafer MJ et al. Cellular senescence mediates fibrotic pulmonary disease. Nat. Commun 8, 14532 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Takahashi K & Yamanaka S Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 (2006). [DOI] [PubMed] [Google Scholar]
  • 93.Israel MA et al. Probing sporadic and familial Alzheimer’s disease using induced pluripotent stem cells. Nature 482, 216–220 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Liu GH et al. Recapitulation of premature ageing with iPSCs from Hutchinson–Gilford progeria syndrome. Nature 472, 221–225 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Miller JD et al. Human iPSC-based modeling of late-onset disease via progerin-induced aging. Cell Stem Cell 13, 691–705 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Rando TA & Chang HY Aging, rejuvenation, and epigenetic reprogramming: resetting the aging clock. Cell 148, 46–57 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Lapasset L et al. Rejuvenating senescent and centenarian human cells by reprogramming through the pluripotent state. Genes Dev. 25, 2248–2253 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Suhr ST et al. Mitochondrial rejuvenation after induced pluripotency. PLoS ONE 5, e14095 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Mertens J et al. Directly reprogrammed human neurons retain aging-associated transcriptomic signatures and reveal age-related nucleocytoplasmic defects. Cell Stem Cell 17, 705–718 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Lo Sardo V et al. Influence of donor age on induced pluripotent stem cells. Nat. Biotechnol 35, 69–74 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Ohnishi K et al. Premature termination of reprogramming in vivo leads to cancer development through altered epigenetic regulation. Cell 156, 663–677 (2014). [DOI] [PubMed] [Google Scholar]
  • 102.Mosteiro L et al. Tissue damage and senescence provide critical signals for cellular reprogramming in vivo. Science 354, aaf4445 (2016). [DOI] [PubMed] [Google Scholar]
  • 103.Abad M et al. Reprogramming in vivo produces teratomas and iPS cells with totipotency features. Nature 502, 340–345 (2013). [DOI] [PubMed] [Google Scholar]
  • 104.Mosteiro L, Pantoja C, de Martino A & Serrano M Senescence promotes in vivo reprogramming through p16INK4a and IL-6. Aging Cell 17, e12711 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Ocampo A et al. In vivo amelioration of age-associated hallmarks by partial reprogramming. Cell 167, 1719–1733.e12 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Falick Michaeli T et al. The rejuvenating effect of pregnancy on muscle regeneration. Aging Cell 14, 698–700 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Polo JM et al. A molecular roadmap of reprogramming somatic cells into iPS cells. Cell 151, 1617–1632 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Papp B & Plath K Epigenetics of reprogramming to induced pluripotency. Cell 152, 1324–1343 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Wang R et al. Rapamycin inhibits the secretory phenotype of senescent cells by a Nrf2-independent mechanism. Aging Cell 16, 564–574 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Iglesias-Bartolome R et al. mTOR inhibition prevents epithelial stem cell senescence and protects from radiation-induced mucositis. Cell Stem Cell 11, 401–414 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Demidenko ZN et al. Rapamycin decelerates cellular senescence. Cell Cycle 8, 1888–1895 (2009). [DOI] [PubMed] [Google Scholar]
  • 112.Soria-Valles C et al. NF-κB activation impairs somatic cell reprogramming in ageing. Nat. Cell Biol 17, 1004–1013 (2015). [DOI] [PubMed] [Google Scholar]
  • 113.Brady JJ et al. Early role for IL-6 signalling during generation of induced pluripotent stem cells revealed by heterokaryon RNA-seq. Nat. Cell Biol 15, 1244–1252 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Lee J et al. Activation of innate immunity is required for efficient nuclear reprogramming. Cell 151, 547–558 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Chiche A et al. Injury-induced senescence enables in vivo reprogramming in skeletal muscle. Cell Stem Cell 20, 407–414.e4 (2017). [DOI] [PubMed] [Google Scholar]
  • 116.Franceschi C et al. Inflamm-aging. An evolutionary perspective on immunosenescence. Ann. NY Acad. Sci 908, 244–254 (2000). [DOI] [PubMed] [Google Scholar]
  • 117.Youm YH et al. Canonical Nlrp3 inflammasome links systemic low-grade inflammation to functional decline in aging. Cell Metab. 18, 519–532 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Strong R et al. Nordihydroguaiaretic acid and aspirin increase lifespan of genetically heterogeneous male mice. Aging Cell 7, 641–650 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Hundal RS et al. Mechanism by which high-dose aspirin improves glucose metabolism in type 2 diabetes. J. Clin. Invest 109, 1321–1326 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Gasparini L, Ongini E & Wenk G Non-steroidal anti-inflammatory drugs (NSAIDs) in Alzheimer’s disease: old and new mechanisms of action. J. Neurochem 91, 521–536 (2004). [DOI] [PubMed] [Google Scholar]
  • 121.Wan QL, Zheng SQ, Wu GS & Luo HR Aspirin extends the lifespan of Caenorhabditis elegans via AMPK and DAF-16/FOXO in dietary restriction pathway. Exp. Gerontol 48, 499–506 (2013). [DOI] [PubMed] [Google Scholar]
  • 122.Zhang G et al. Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature 497, 211–216 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Liu M et al. Resveratrol inhibits mTOR signaling by promoting the interaction between mTOR and DEPTOR. J. Biol. Chem 285, 36387–36394 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Conboy IM, Conboy MJ & Rebo J Systemic problems: a perspective on stem cell aging and rejuvenation. Aging 7, 754–765 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Tran D et al. Insulin-like growth factor-1 regulates the SIRT1–p53 pathway in cellular senescence. Aging Cell 13, 669–678 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Laberge RM et al. MTOR regulates the pro-tumorigenic senescence-associated secretory phenotype by promoting IL1A translation. Nat. Cell Biol 17, 1049–1061 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Herranz N et al. mTOR regulates MAPKAPK2 translation to control the senescence-associated secretory phenotype. Nat. Cell Biol 17, 1205–1217 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Han X et al. AMPK activation protects cells from oxidative stress-induced senescence via autophagic flux restoration and intracellular NAD+ elevation. Aging Cell 15, 416–427 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Wu Y et al. Autophagy and mTORC1 regulate the stochastic phase of somatic cell reprogramming. Nat. Cell Biol 17, 715–725 (2015). [DOI] [PubMed] [Google Scholar]
  • 130.Chen T et al. Rapamycin and other longevity-promoting compounds enhance the generation of mouse induced pluripotent stem cells. Aging Cell 10, 908–911 (2011). [DOI] [PubMed] [Google Scholar]
  • 131.Benayoun BA, Pollina EA & Brunet A Epigenetic regulation of ageing: linking environmental inputs to genomic stability. Nat. Rev. Mol. Cell Biol 16, 593–610 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Horvath S DNA methylation age of human tissues and cell types. Genome Biol. 14, R115 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Hannum G et al. Genome-wide methylation profiles reveal quantitative views of human aging rates. Mol. Cell 49, 359–367 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Narita M et al. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell 113, 703–716 (2003). [DOI] [PubMed] [Google Scholar]
  • 135.Chandra T et al. Independence of repressive histone marks and chromatin compaction during senescent heterochromatic layer formation. Mol. Cell 47, 203–214 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Foran E et al. Upregulation of DNA methyltransferase-mediated gene silencing, anchorage-independent growth, and migration of colon cancer cells by interleukin-6. Mol. Cancer Res 8, 471–481 (2010). [DOI] [PubMed] [Google Scholar]
  • 137.Hodge DR et al. Interleukin-6 regulation of the human DNA methyltransferase (HDNMT) gene in human erythroleukemia cells. J. Biol. Chem 276, 39508–39511 (2001). [DOI] [PubMed] [Google Scholar]
  • 138.Kim CH et al. Short-term calorie restriction ameliorates genomewide, age-related alterations in DNA methylation. Aging Cell 15, 1074–1081 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Hahn O et al. Dietary restriction protects from age-associated DNA methylation and induces epigenetic reprogramming of lipid metabolism. Genome Biol. 18, 56 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Rubinsztein DC, Marino G & Kroemer G Autophagy and aging. Cell 146, 682–695 (2011). [DOI] [PubMed] [Google Scholar]
  • 141.Barzilai N, Crandall JP, Kritchevsky SB & Espeland MA Metformin as a tool to target aging. Cell Metab. 23, 1060–1065 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Kang C et al. The DNA damage response induces inflammation and senescence by inhibiting autophagy of GATA4. Science 349, aaa5612 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Gewirtz DA Autophagy and senescence: a partnership in search of definition. Autophagy 9, 808–812 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Ho TT et al. Autophagy maintains the metabolism and function of young and old stem cells. Nature 543, 205–210 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Garcia-Prat L et al. Autophagy maintains stemness by preventing senescence. Nature 529, 37–42 (2016). [DOI] [PubMed] [Google Scholar]
  • 146.Leeman DS et al. Lysosome activation clears aggregates and enhances quiescent neural stem cell activation during aging. Science 359, 1277–1283 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Prigione A et al. Mitochondrial-associated cell death mechanisms are reset to an embryonic-like state in aged donor-derived iPS cells harboring chromosomal aberrations. PLoS ONE 6, e27352 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Passos JF et al. Feedback between p21 and reactive oxygen production is necessary for cell senescence. Mol. Syst. Biol 6, 347 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Passos JF et al. Mitochondrial dysfunction accounts for the stochastic heterogeneity in telomere-dependent senescence. PLoS Biol. 5, e110 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Wiley CD et al. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 23, 303–314 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Lakowski B & Hekimi S Determination of life-span in Caenorhabditis elegans by four clock genes. Science 272, 1010–1013 (1996). [DOI] [PubMed] [Google Scholar]
  • 152.Liu X et al. Evolutionary conservation of the clk-1-dependent mechanism of longevity: loss of mclk1 increases cellular fitness and lifespan in mice. Genes Dev. 19, 2424–2434 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Dillin A et al. Rates of behavior and aging specified by mitochondrial function during development. Science 298, 2398–2401 (2002). [DOI] [PubMed] [Google Scholar]
  • 154.Owusu-Ansah E, Song W & Perrimon N Muscle mitohormesis promotes longevity via systemic repression of insulin signaling. Cell 155, 699–712 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Copeland JM et al. Extension of Drosophila life span by RNAi of the mitochondrial respiratory chain. Curr. Biol 19, 1591–1598 (2009). [DOI] [PubMed] [Google Scholar]
  • 156.Fatt M et al. Metformin acts on two different molecular pathways to enhance adult neural precursor proliferation/self-renewal and differentiation. Stem Cell Rep. 5, 988–995 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Beerman I, Seita J, Inlay MA, Weissman IL & Rossi DJ Quiescent hematopoietic stem cells accumulate DNA damage during aging that is repaired upon entry into cell cycle. Cell Stem Cell 15, 37–50 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Poulos MG et al. Endothelial transplantation rejuvenates aged hematopoietic stem cell function. J. Clin. Invest 127, 4163–4178 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Sepulveda JC et al. Cell senescence abrogates the therapeutic potential of human mesenchymal stem cells in the lethal endotoxemia model. Stem Cells 32, 1865–1877 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Fontana L et al. The effects of graded caloric restriction: XII. Comparison of mouse to human impact on cellular senescence in the colon. Aging Cell 17, e12746 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Banito A et al. Senescence impairs successful reprogramming to pluripotent stem cells. Genes Dev. 23, 2134–2139 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Li H et al. The Ink4/Arf locus is a barrier for iPS cell reprogramming. Nature 460, 1136–1139 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Ritschka B et al. The senescence-associated secretory phenotype induces cellular plasticity and tissue regeneration. Genes Dev. 31, 172–183 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Ingram DK & de Cabo R Calorie restriction in rodents: caveats to consider. Ageing Res. Rev 39, 15–28 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Mannick JB et al. mTOR inhibition improves immune function in the elderly. Sci. Transl Med 6, 268ra179 (2014). [DOI] [PubMed] [Google Scholar]
  • 166.Mak SS, Moriyama M, Nishioka E, Osawa M & Nishikawa S Indispensable role of Bcl2 in the development of the melanocyte stem cell. Dev. Biol 291, 144–153 (2006). [DOI] [PubMed] [Google Scholar]
  • 167.McDonnell TJ et al. bcl-2-Immunoglobulin transgenic mice demonstrate extended B cell survival and follicular lymphoproliferation. Cell 57, 79–88 (1989). [DOI] [PubMed] [Google Scholar]
  • 168.Zhang H et al. Bcl-2 family proteins are essential for platelet survival. Cell Death Differ. 14, 943–951 (2007). [DOI] [PubMed] [Google Scholar]
  • 169.Bernardes de Jesus B et al. Telomerase gene therapy in adult and old mice delays aging and increases longevity without increasing cancer. EMBO Mol. Med 4, 691–704 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Matheu A et al. Anti-aging activity of the Ink4/Arf locus. Aging Cell 8, 152–161 (2009). [DOI] [PubMed] [Google Scholar]
  • 171.Gonzalez-Navarro H et al. Increased dosage of Ink4/Arf protects against glucose intolerance and insulin resistance associated with aging. Aging Cell 12, 102–111 (2013). [DOI] [PubMed] [Google Scholar]
  • 172.Carrasco-Garcia E, Arrizabalaga O, Serrano M, Lovell-Badge R & Matheu A Increased gene dosage of Ink4/Arf and p53 delays age-associated central nervous system functional decline. Aging Cell 14, 710–714 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.van Praag H, Shubert T, Zhao C & Gage FH Exercise enhances learning and hippocampal neurogenesis in aged mice. J. Neurosci 25, 8680–8685 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Luo L et al. Chronic resistance training activates autophagy and reduces apoptosis of muscle cells by modulating IGF-1 and its receptors, Akt/mTOR and Akt/FOXO3a signaling in aged rats. Exp. Gerontol 48, 427–436 (2013). [DOI] [PubMed] [Google Scholar]
  • 175.Valdez G et al. Attenuation of age-related changes in mouse neuromuscular synapses by caloric restriction and exercise. Proc. Natl Acad. Sci. USA 107, 14863–14868 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Valenzano DR, Terzibasi E, Cattaneo A, Domenici L & Cellerino A Temperature affects longevity and age-related locomotor and cognitive decay in the short-lived fish Nothobranchius furzeri. Aging Cell 5, 275–278 (2006). [DOI] [PubMed] [Google Scholar]
  • 177.Zhang B et al. Environmental temperature differentially modulates C. elegans longevity through a thermosensitive TRP channel. Cell Rep. 11, 1414–1424 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Lee SJ & Kenyon C Regulation of the longevity response to temperature by thermosensory neurons in Caenorhabditis elegans. Curr. Biol 19, 715–722 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Conti B et al. Transgenic mice with a reduced core body temperature have an increased life span. Science 314, 825–828 (2006). [DOI] [PubMed] [Google Scholar]
  • 180.Smith P et al. Regulation of life span by the gut microbiota in the short-lived African turquoise killifish. eLife 6, e27014 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Zhang Y et al. Hypothalamic stem cells control ageing speed partly through exosomal miRNAs. Nature 548, 52–57 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Wei M et al. Fasting-mimicking diet and markers/risk factors for aging, diabetes, cancer, and cardiovascular disease. Sci. Transl Med 9, eaai8700 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Singh M et al. Effect of low-dose rapamycin on senescence markers and physical functioning in older adults with coronary artery disease: results of a pilot study. J. Frailty Aging 5, 204–207 (2016). [DOI] [PubMed] [Google Scholar]
  • 184.Gandini S et al. Metformin and cancer risk and mortality: a systematic review and meta-analysis taking into account biases and confounders. Cancer Prev. Res 7, 867–885 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Holman RR, Paul SK, Bethel MA, Matthews DR & Neil HA 10-year follow-up of intensive glucose control in type 2 diabetes. N. Engl. J. Med 359, 1577–1589 (2008). [DOI] [PubMed] [Google Scholar]
  • 186.Yosef R et al. Directed elimination of senescent cells by inhibition of BCL-W and BCL-XL. Nat. Commun 7, 11190 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Abbott A Infusions of young blood tested in patients with dementia. Nature News (1 November 2017). [Google Scholar]
  • 188.Mahmoudi S & Brunet A Bursts of reprogramming: a path to extend lifespan? Cell 167, 1672–1674 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Table 1

RESOURCES