Blog

Cellular senescence in ageing: from mechanisms to therapeutic opportunities

Cellular senescence in ageing: from mechanisms to therapeutic opportunities

Nat Rev Mol Cell Biol. Author manuscript; available in PMC 2021 Aug 6.

Published in final edited form as:

Nat Rev Mol Cell Biol. 2021 Feb; 22(2): 75–95.

Published online 2020 Dec 16. doi: 10.1038/s41580-020-00314-w

PMCID: PMC8344376

NIHMSID: NIHMS1728344

PMID: 33328614

Cellular senescence in ageing: from mechanisms to therapeutic opportunities

Raffaella Di Micco,1, Valery Krizhanovsky,2 Darren Baker,3,4 and Fabrizio d’Adda di Fagagna5,6,

Author information Copyright and License information Disclaimer

The publisher's final edited version of this article is available at Nat Rev Mol Cell Biol

 

Abstract

Cellular senescence, first described in vitro in 1961, has become a focus for biotech companies that target it to ameliorate a variety of human conditions. Eminently characterized by a permanent proliferation arrest, cellular senescence occurs in response to endogenous and exogenous stresses, including telomere dysfunction, oncogene activation and persistent DNA damage. Cellular senescence can also be a controlled programme occurring in diverse biological processes, including embryonic development. Senescent cell extrinsic activities, broadly related to the activation of a senescence-associated secretory phenotype, amplify the impact of cell-intrinsic proliferative arrest and contribute to impaired tissue regeneration, chronic age-associated diseases and organismal ageing. This Review discusses the mechanisms and modulators of cellular senescence establishment and induction of a senescence-associated secretory phenotype, and provides an overview of cellular senescence as an emerging opportunity to intervene through senolytic and senomorphic therapies in ageing and ageing-associated diseases.

Cellular senescence occurs in response to many different triggers, including DNA damage, telomere dysfunction, oncogene activation and organelle stress, and has been linked to processes such as tumour suppression, tissue repair, embryogenesis and organismal ageing. Hayflick and Moorhead demonstrated in 1961 that normal cultured human fibroblasts display a finite capacity for cell division before entering an irreversible growth arrest known as replicative senescence1. This led to the hypothesis that tissue ageing is caused by cells progressively losing their ability to proliferate, which is essential to replace damaged cells that naturally accumulate over time2. However, it took several decades to develop tools to prove that senescent cell accumulation promotes organismal ageing and dysfunction.

The first hurdle involved identifying selective markers to detect these cells in living tissues. Senescence-associated-β-galactosidase (SA-β-gal) activity, a simple colorimetric assay, was one of the first biomarkers described, and became instrumental in demonstrating that cells with features of senescence accumulate at sites of ageing-associated diseases and in aged tissues in a variety of mammals. A distinctive feature of senescent cells is the increased expression of cell cycle-inhibitory proteins, collectively known as cyclin-dependent kinase inhibitors. The cyclin-dependent kinase inhibitor with the most prominent role in senescent cell accumulation during ageing is p16INK4A — hereafter referred to as p16 (REFS3,4) — as it is crucial for durably maintaining the state of proliferative arrest. Indeed, mice lacking p16 are predisposed to spontaneous tumour formation5. In the late 1990s, it was found that cellular senescence is prematurely induced by excessive oncogenic signalling or loss of tumour suppression6. Senescence induction was later shown to be caused by aberrant DNA replication and DNA damage accumulation7,8, thus restricting the proliferation of damaged precancerous cells. However, none of these features is universal for senescent cells, and it is important to test several biomarkers simultaneously to define the senescence state.

To reconcile the seemingly opposing pro-ageing and anticancer roles of senescent cells from an evolutionary perspective, senescent cells were proposed to fit the antagonistic pleiotropy theory of ageing, which posits that natural selection favours genes that promote reproductive fitness early in life, which may be accompanied by unselected consequences with negative effects later in life9, although this has not been proven10. Alternatively, it is conceivable that evolutionary cost and benefit theory is relevant for senescence. This theory implies that senescent cells have beneficial effects throughout life (for example, limiting tissue damage and suppressing tumorigenesis), but the cost of these effects overcomes the benefits in old age. Several approaches have recently enabled the establishment of a causative role for senescent cells in many diseases. These include the development of INK-ATTAC11 and p16-3Mr12 transgenic mouse models in which p16-expressing cells can be selectively eliminated, and of senolytic and senomorphic pharmacological agents. Senolytics target senescent cells for elimination, whereas senomorphics modulate the properties of senescent cells without eliminating them13.

INK-ATTAC

Transgenic mouse model with drug-inducible caspase 8 under the control of a minimal p16 promoter element active in senescent cells to allow selective elimination of p16-expressing senescent cells.

p16-3MR

Transgenic mouse model expressing a trimodal reporter of red fluorescent protein, luciferase and herpes simplex virus thymidine kinase under the control of the p16 promoter to allow tracking and elimination of p16-expressing senescent cells.

In this Review, we first describe the properties of senescent cells and the mechanisms that promote this phenotype. We then discuss the implication of senescent cells in diverse biological processes, and how their removal or the attenuation of their properties could be exploited for therapeutic intervention and to increase healthspan.

 

Inducers and features of cellular senescence

Cellular senescence is a stable and terminal state of growth arrest in which cells are unable to proliferate despite optimal growth conditions and mitogenic stimuli (BOXES 1,2FIG. 1). Senescent cells have increased resistance to apoptotic cell death owing to upregulation of cell survival pathways, including the BCL-2 family of antiapoptotic proteins, even on exogenous stress exposure14,15. Whether this prolonged viability is the result of selection for the most death-resistant cells or whether it is an intrinsic property of the senescence programme, especially in vivo, remains to be established. The molecular mechanisms that determine the choice between apoptosis and senescence remain unclear, but it is possible that cell fate depends on the intensity and duration of the initial stimulus, as well as the nature of the damage and the cell type16. Because senescence and apoptosis programmes converge on crucial components, including activation of the p53 pathway, it is also possible that senescent cell resistance to apoptosis depends on p53 levels and activity17. Although senescence was believed to be a permanent condition of cell cycle arrest, recent evidence indicates that, at least in the context of tumour formation and anticancer therapies, the establishment of cellular senescence might involve epigenetic mechanisms that reprogramme cancer cells towards a certain degree of stemness in a cell-autonomous fashion18. Of note, the establishment of senescence is a dynamic process, whereby overlapping but distinct molecular pathways are engaged at different stages, from immediately on cell cycle exit to late, senescence stages19.

 

 

Fig. 1 |

Senescence drivers and phenotypes.

Nuclear DNA damage is often causatively associated with senescence establishment. DNA damage activates a signalling cascade defined as DNA damage response (DDR), characterized by phosphorylated histone H2AX (γH2AX), 53BP1 and MDC1, the apical kinases ataxia telangiectasia mutated (ATM) and ATR and the downstream kinases CHK2 and CHK1. Signals ultimately converge on p53 activation, which in turn elicits cell cycle arrest. Prolonged DDR activation triggers senescence. One or a few DDR signalling telomeres, the ends of chromosomes, are sufficient to trigger replicative cell senescence. Oncogene activation is also a powerful senescence trigger. Specifically, most activated oncogenes, partly via reactive oxygen species (ROS) production, induce hyperproliferation and altered DNA replication patterns that ultimately result in replication stress and DNA damage accumulation at fragile sites, which include telomeres. Besides prolonged DDR activation, senescence features include cell cycle arrest (by upregulation of p21 and p16 cell cycle inhibitors), oxidative damage (as detected by increased ROS levels), upregulation of the BCL-2 family of antiapoptotic proteins, which induce resistance to apoptosis, metabolic changes (including senescence-associated-β-galactosidase (SA-β-gal) accumulation), senescence-associated heterochromatin foci (SAHF) and a senescence-associated secretory phenotype (SASP).

Box 1 |

Senescence biomarkers

A major limitation in the senescence field is the lack of single, universal or model-specific biomarkers to identify senescent cells in culture or tissue samples. At present, the identification of senescent cells relies on a combination of multiple markers that, when present simultaneously, can discriminate between stably arrested senescent cells and quiescent or differentiated counterparts.

The first and still the most widely used biomarker to detect senescent cells in cultured cells and in fresh tissue samples is the accumulation of a lysosomal enzyme termed ‘senescence-associated-β-galactosidase’ (SA-β-gal)275. This marker is detectable by histochemical staining in most senescent cells and is generally not found in presenescent, quiescent or immortal and transformed cells, although SA-β-gal can also accumulate in serum-starved or overconfluent cells in tissue culture and may mark a specific subpopulation of macrophages in vivo as part of a reversible response to immune stimuli276. Lipofuscin accumulation is another feature of senescent cells. A recently developed method based on biotin-linked Sudan black B analogue is emerging as a reliable detection system to trace senescent cells in a variety of cell and tissue types277.

Another peculiarity of senescent cells is the abnormally enlarged and flat morphology with disproportionate increase in the cytoplasm-to-nucleus ratio. While this bulky cytoplasm was originally described as a feature accompanying the establishment of cell senescence, a recent study suggests that increased cell size may play a causative role in driving the senescence-associated growth arrest278. In addition, in vivo, SA-β-gal-positive senescent cells have increased cell size compared with SA-β-gal-negative cells as identified on a single-cell level279. Another obvious marker for senescent cells is the lack of DNA replication, which is typically detected by the incorporation of nucleoside analogous (for example, 5-bromodeoxyuridine or [3H]thymidine) or by immunostaining for proliferation markers, such as PCNA and Ki-67. These markers do not distinguish between senescent cells and quiescent or differentiated postmitotic cells.

p21 and p16 are two cyclin-dependent kinase inhibitors that are components of the tumour suppressor pathways governed by p53 and RB, and often accumulate in senescent cells. Because p21 and p16 expression levels are sufficient to establish and maintain the senescence-associated growth arrest, they are used to identify senescent cells in tissues and cultured cells. p16, in particular, was used as a surrogate senescence marker for the generation of engineered mouse models where selective eradication of senescent cells has been tested. However, not all senescent cell types express p16 as it can be expressed also by some tumour cells, especially those that have lost RB functions280.

Nuclear senescence-associated heterochromatin foci (SAHF) are also used to identify senescent cells, but they appear to be specific to the senescence programme induced by activated oncogenes and DNA replication stressors52. Persistent DNA damage response factors accumulating at sites of damage as cytologically detectable nuclear foci are also used as markers of senescent cells, and when accumulating at telomeric sequences, telomere-associated foci represent a robust marker of the senescent state24,35,281.

Lastly, components of the senescence-associated secretory phenotype (SASP), mainly the proinflammatory cytokines interleukin-6 (Il-6) and Il-8, may be used at the transcript and protein levels to evaluate general tissue or cell culture senescence. However, SASP alone cannot be used as a reliable senescence biomarker: indeed, senescence triggered by p16 overexpression does not entail an altered SASP transcriptional programme85. Overall, the search for universal senescence biomarkers is constantly challenged by the evidence that senescence phenotypes are highly heterogeneous and may differ depending on the initial trigger and the cell type under study. Therefore, transcriptomic and proteomic studies up to the single-cell level in relevant cell and tissue types will be of paramount importance to find unique or common markers of the senescence state, including cell surface molecules that will allow the prospective isolation of senescent cells from aged and diseased tissues. More recently, the development of innovative imaging-based tools and fluorescent tracers to monitor in real time senescence burden and monitor the therapeutic activity of senotherapies in clinical samples gained the interest of the scientific community282284 and may represent a turning point for senescence-based translational medicine applications.

Box 2 |

Senescence and autophagy

Dysfunctional cellular organelles, such as mitochondria and lysosomes, are usually degraded through the activation of an intracellular degradation system named ‘autophagy’285. However, whether autophagy promotes senescence induction or is an alternative prosurvival mechanism lost during ageing is a matter of intense scientific investigation. Indeed, it was reported that a selective autophagy pathway, via mammalian target of rapamycin (mTOR) activation, contributes to sustain the protein synthesis of many senescence-associated secretory phenotype factors mainly in oncogene-induced senescent cells and that downregulation of several autophagy regulators delays the establishment of oncogene-induced senescence286,287. more recently, LC3B, a ubiquitin-like autophagic protein, was found to be associated with the nuclear envelope protein lamin B1 and to contribute to its degradation in the lysosomes in oncogene-induced senescent cells59,60. Importantly, lamin-associated chromatin domains are also transported out from the nucleus to the lysosomes via the same mechanism and contribute to the presence of cytosolic chromatin fragments that accumulate in senescent cells. Inhibition of autophagy prevents lamin B1 degradation and ensures nuclear envelope integrity in senescent cells59.

In the context of therapy-induced senescence for cancer cells, it was also shown that autophagy was triggered as a consequence of senescence establishment to cope with the increased load of accumulating toxic macromolecules, and its pharmacological targeting led to senescent cell elimination288. However, autophagy has also been considered to suppress senescence by promoting the degradation of damaged organelles and other cellular components, and several studies have supported such a view. In adult muscle stem cells, basal autophagy maintains stemness by repressing senescence. During ageing, the autophagic activity in muscle stem cells (satellite cells) declines alongside stem cell regenerative capacity, with consequent accumulation of senescent satellite cells in geriatric mice. autophagy restoration in old satellite cells prevents senescence and rescues their regenerative capacity289. Similarly, autophagy protects against oxidative stress-induced senescence. enhancing autophagic activity under excessive oxidative stress by mTOR inhibition delays cellular senescence and functionally restores both mitochondrial and lysosomal functions290. Further supporting a role of autophagy in preventing senescence, a recent high-throughput screening aimed at identifying compounds that alleviate replicative senescence revealed that the ataxia telangiectasia mutated (ATM) inhibitor KU-60019 enhances the autophagic flux by restoring functional lysosomal activity by blocking the phosphorylation of the vacuolar protein ATP6V1G1. ATM inhibitor treatment also recovered mitochondrial functions and alleviated senescence phenotypes274. Together, these seemingly opposite roles may reflect a complex and reciprocal regulation of autophagy and cell senescence that can be linked to several senescence triggers, distinct cell types and a unique spatio-temporal activation of the autophagic programme acting in the senescence regulatory network.

Cellular senescence and the DNA damage response.

Several stressors can induce cellular senescence. Nuclear DNA damage is often reported as a commonly underlying cause of senescence, mainly in the form of DNA double-strand breaks (DSBs)20 that activate the DNA damage response (DDR) pathway (FIG. 1). DDR exerts checkpoint functions to block cell cycle progression and prevent the propagation of corrupted genetic information to daughter cells. Some DDR factors accumulate at sites of DNA damage and form cytologically detectable nuclear foci composed of extended chromatin modification events, such as the phosphorylation of histone H2AX, and the proteins associated with them, including MDC1, 53BP1 and the activated form of the kinase ataxia telangiectasia mutated (ATM)21. These foci mark individual sites of DNA damage and contribute to checkpoint enforcement and cell cycle arrest, until damage has been repaired. If DNA damage persists, it causes prolonged DDR signalling and protracted proliferative arrest in the form of cellular senescence22. The recent demonstration that persistent DDR foci observed in cultured senescent cells contain unrepaired DSBs23 supports the notion that cellular senescence is akin to prolonged checkpoint activation. Inhibition of DDR signalling kinases (ATM, ATR, CHK1 and CHK2) allows senescent cells to re-enter the cell cycle7,24,25. At the bottom of the DDR cascade, the tumour suppressor p53, which is a target of ATM and its paralogue ATR, is activated and stimulates the expression of the cyclin-dependent kinase inhibitor p21, an essential mediator of senescence-associated cell cycle arrest. p16, an inhibitor of CDK4 and CDK6, is also key in several types of senescence26; p21 is activated early on senescence entry and p16 is activated later, probably to maintain the senescence phenotype27. In addition to the DDR cascade being activated, the tumour suppressor ARF stabilizes p53, which contributes to the induction of senescence28. Efforts have been devoted to assessing the contribution of these two major routes, the DDR and ARF pathways, to p53-dependent senescence establishment especially in response to oncogenic challenges. The original view, based mainly on murine studies, was that the DDR and ARF play antagonistic roles, as ARF was transcriptionally activated during tumorigenesis in a DDR-independent manner29,30. More recently, a tight regulatory network in human cancer models was reported, whereby ATM suppresses ARF levels and ARF acts as a secondary barrier to cancer progression when ATM is inactivated31. Consistent with this temporal regulation, DDR precedes ARF engagement, whose activation is detected at later stages of cancer progression and less frequently than DDR.

Telomere shortening and damage.

One of the first and best characterized mechanisms of cellular senescence induction is telomere shortening. As the standard DNA replication apparatus is unable to fully duplicate chromosomal DNA ends, in the absence of telomere maintenance mechanisms such as the expression of telomerase or recombination among telomeres, telomeres shorten with each round of DNA replication. Below a certain length, the loss of telomere-capping factors or protective structures makes critically short telomeres resemble one-ended DSBs and thus triggers a DDR that is very similar to that triggered by DNA DSBs24,32 (FIG. 1). One or a few DDR signalling telomeres are sufficient to trigger replicative cellular senescence33, and forced expression of telomerase prevents cellular senescence and promotes unlimited cell proliferation34.

Persistent DDR activation occurs also at telomeres that are not critically shortened, in non-dividing cells exposed to exogenous genotoxic treatments and in non-dividing ageing cells, because repair is much less efficient when DSBs are localized within telomeres3537. As telomeric DSBs persist, cellular senescence is established and maintained. Thus, persistent DDR activation at telomeres, which is a trigger of cellular senescence, can occur both upon telomere shortening in proliferating cells and upon telomeric DNA damage also in non-proliferating (quiescent or terminally differentiated) cells, independently of telomere length38.

Oncogene-induced senescence.

Oncogene activation is a powerful inducer of cellular senescence. Oncogene expression triggers an initial hyperproliferative phase that is intrinsically associated with altered DNA replication, which eventually engages DDR pathways and causes senescence7,8,39. This process is known as oncogene-induced senescence (OIS). Loss of tumour suppressor expression can also induce proliferation arrest, as exemplified by PTEN loss-induced cellular senescence (PICS). Although initially PTEN loss-induced cellular senescence was not associated with DDR activation40, it was later found to be associated with hyperproliferation, DDR engagement and cellular senescence in vivo41. Noteworthy, unlike oncogenic RAS or BRAF, activation of the PI3K–AKT pathway promotes p53-dependent senescence often in the absence of detectable hyperproliferation and strong DNA damage accumulation42,43, suggesting distinct underlying mechanisms.

Telomeres are hypersensitive to DNA replication stress, including that induced by oncogenes and accumulation of oncogene-induced telomeric dysfunction, and a DDR has been observed in hyperplastic cancer lesions in humans44. Reactive oxygen species (ROS) accumulate in tumours, and in this context, in addition to their recognized role as DNA-damaging agents, they can act as signalling molecules that mediate pro-mitogenic oncogene functions. Recently, this paradoxical role of ROS in promoting cell proliferation and senescence-associated DNA damage was partly solved by the unexpected discovery that oncogene-induced ROS, generated by NADPH oxidases, can induce cellular senescence by boosting the initial hyperproliferative phase associated with altered DNA replication and DNA damage accumulation45 (FIG. 1).

Mitochondrial dysfunctions and cellular senescence.

Increased oxidative stress in senescent cells has been linked to the accumulation of dysfunctional mitochondria. Indeed, senescent cells are characterized by changes in mitochondrial mass, membrane potential and mitochondrial morphology46. Dysfunctional mitochondria may play an important role in senescence establishment, as depletion of mitochondrial sirtuins, a group of evolutionarily conserved proteins that regulate ageing across different species, as well as selective chemical inhibition of mitochondrial function, triggers senescence47. There is evidence in support of a reciprocal influence between nuclear DNA damage and mitochondrial dysfunction48. Of note, mitochondrial dysfunction-associated senescence (MiDAs), which is characterized by a distinct phenotype, exhibits a unique cell-non-autonomous programme that is potentially responsible for the altered metabolism and aberrant adipocyte differentiation observed in aged animals47.

Sirtuins

Nicotinamide dinucleotide (NAD+)-dependent deacylases that regulate diverse cellular processes, including DNA repair, inflammation, metabolism and ageing.

Mitochondrial dysfunction-associated senescence

(MiDAs). Mitochondrial damage triggers senescence with a distinct secretory phenotype that lacks IL-1-dependent inflammation.

Chromatin changes in senescent cells.

Most senescent cells display profound changes in the epigenome and chromatin organization. These changes have been linked to both the cell-autonomous and paracrine aspects (that is, the effect on surrounding cells) of senescence-associated proliferation arrest. Senescence-associated heterochromatin foci (SAHF) are spatially organized heterochromatic domains that can be detected as dense 4′,6-diamidino-2-phenylindole (DAPI)-positive nuclear structures that are enriched in repressive chromatin marks and proteins, including trimethylated histone H3 Lys9 (H3K9me3), heterochromatin protein 1 (HP1), high mobility group protein A (HMGA) factors, histone variant macroH2A and histone co-chaperones HIRA and ASF1A4951. However, SAHF are not universal markers of senescence; they are most robustly observed upon oncogene activation and form in a DNA replication- and ATR-dependent manner52. SAHF were originally proposed to repress genes promoting cell cycle progression49,53,54. Rather, SAHF enforce a DDR-resistant heterochromatin structure that restrains DDR signalling52. Indeed, treatment with histone deacetylase (HDAC) inhibitors, which induce chromatin relaxation, boosts DDR signalling with consequent cell death by apoptosis. This treatment is probably the first reported example of a successful senolytic approach52, which was later supported by the reported senolytic activity of the HDAC inhibitor panobinostat55. HDAC inhibitors can also induce cellular senescence in normal human fibroblasts, which may be related to their impact on the DDR56.

Another chromatin feature of senescent cells is the unfolding of constitutive heterochromatin domains characterized mainly by distension of pericentromeric satellite sequences, which was observed in different species and following different modes of senescence induction57. These changes in chromatin structure were not linked to the selective removal of repressive histone marks, but were associated with changes in nuclear structural proteins, including breakdown of the nuclear lamina57. Loss of nuclear lamina can lead to the release of cytosolic chromatin fragments (CCFs) in the cytoplasm of senescent cells5860. Although yet unprobed, in the context of oncogene-induced DNA replication, hard-to-replicate genomic regions such as fragile sites61, telomeric sequences44 and repetitive DNA62 probably contribute to CCFs. It remains unclear whether CCFs are formed only in deeply senescent cells. Importantly, CCFs dictate senescence-associated paracrine functions through the activation of the cyclic GMP–AMP synthase (cGAS) and the adaptor stimulator of interferon genes (STING) pathway (discussed later). Low doses of HDAC inhibitors have been reported to reduce CCFs and suppress the senescence-associated secretory phenotype (SASP)63.

Recent technological advances in genome-wide mapping of chromatin modifications led to the generation of a molecular blueprint of senescence establishment and maintenance. During replicative senescence, late-replicating, gene-poor regions display widespread DNA hypomethylation, whereas focal hypermethylation is seen at tumour suppressor genes. These observations led to the hypothesis that senescent cells may be epigenetically primed for malignant transformation64. But this hypothesis was recently challenged by the observation that cells with OIS display only limited changes in methylation patterns compared with cells that have bypassed OIS, indicating that tumour-associated methy-lome changes may arise stochastically and independently of the senescence state65. In contrast to what would be observed in the case of DNA methylation changes, oncogene-induced senescent cells and late replicatively senescent fibroblasts exhibit a marked increase in chromatin accessibility at the nucleosomal level, with most of the open chromatin regions mapping to regulatory elements and repeats66,67. Chromatin loosening at genomic repeats results in increased expression levels of transposable elements, which are normally epigenetically silenced and dormant in unstressed cells67. Despite the well-accepted role of transposable elements in triggering genomic instability via transposition, the reactivation of transposable elements also contributes to mediate the non-cell-autonomous functions of senescent cells as detailed later62. Genome-wide analysis of H3 Lys4 trimethylation, H3 Lys27 trimethylation and H3 Lys27 acetylation in senescent cells has also revealed the dynamic acquisition and depletion of large-scale chromatin domains that have been proposed to regulate the expression of key senescence downstream effectors6872.

Go to:

SASP composition and regulation

One potential mechanism through which senescent cells exert their pleiotropic biological functions is the transcriptional activation of a SASP programme characterized by cytokines, chemokines, growth factors and extracellular matrix (ECM) proteases, which may self-reinforce senescence or affect the local tissue microenvironment of senescent cells and possibly the entire organism (FIG. 2). SASP activation is a dynamic process that accompanies senescence establishment. SASP was originally defined as a robust secretory programme comprising dozens if not hundreds of bioactive factors7376.

 

 

Fig. 2 |

SASP regulation.

Senescence-associated secretory phenotype (SASP) activation is a dynamic process that accompanies cell cycle exit initiated by senescence triggers. A core SASP programme comprises mainly proinflammatory interleukin-6 (IL-6), IL-8 and monocyte chemoattractant protein 1 (MCP1), regulated in an IL-1-dependent manner, and enzymes involved in extracellular matrix (ECM) remodelling, such as matrix metalloproteinases (MMPs), serine/cysteine proteinase inhibitors (SERPINs) and tissue inhibitors of metalloproteinases (TIMPs). More recently, additional core SASP effectors released as soluble molecules or in exosomes were identified, including GDF15, STC1 and MMP1. DNA damage response factors, including the upstream DNA damage response kinase induce SASP genes via nuclear factor-κB (NF-κB). The mitogen-activated protein kinase p38 also induces SASP genes by increasing the activity of NF-κB. Activation of several transcription factors and chromatin regulators has been implicated in SASP activation and regulation. NF-κB the transcription factor CCAAT/enhancer-binding protein-β (C/EBPβ) bind promoters of SASP genes and regulate their activation. GATA4 regulates NF-κB and SASP genes indirectly via IL-1 production. The mammalian target of rapamycin (mTOR) pathway also promotes SASP production through increased translation of subsets of mRNAs, including that encoding for IL-1α. In concert with transcription factors, the epigenetic reader bromodomain-containing protein 4 (BRD4), an acetylated histone-binding protein involved in oncogenesis, is recruited to superenhancers adjacent to SASP genes, thus contributing to the proper execution of cellular senescence. BRD4 binds acetylated histone H3 Lys27 (H3K27), thus competing with Polycomb repressor complex 2 (PRC2), which methylates the same histone residue (to give trimethylated H3K27) for transcriptional repression. Consistent with this, PRC2 inhibits SASP genes in senescent cells. More recently, the DNA sensor cyclic GMP–AMP synthase (cGAS) and the adaptor stimulator of interferon genes (STING) have been reported to be major regulators of the SASP programme across species and senescence modes, presumably by activating NF-κB and interferon response factor IRF3 on recognition of cytosol DNA and cytosolic chromatin fragments (CCFs). Aberrant activation of the cGAS–STING pathway could be linked to the downregulation of DNases (for example, DNase 2 and TREX1), enzymes normally involved in cytoplasmic DNA degradation. ATM, ataxia telangiectasia mutated; cGAMP, cyclic GMP–AMP; IL-1R, interleukin-1 receptor.

SASP composition varies depending on the cell type and the nature of the initial stimulus, with the oncogenic trigger greatly amplifying protein secretion compared with replicative or irradiation-induced senescence75. Despite some qualitative and quantitative differences among the SASP in different tissues and senescence models, a core SASP programme comprising mainly proinflammatory interleukin-6 (IL-6), CXC chemokine ligand 8 (CXCL8, hereafter named IL-8) and monocyte chemoattractant protein 1 (MCP1; also known as CCL2) was reported in all types of in vitro-generated senescent cells75. SASP not only includes proinflammatory molecules but also enzymes involved in ECM remodelling, such as matrix metalloproteinases (MMPs)77, serine/cysteine proteinase inhibitors (SERPINs)78 and tissue inhibitors of metalloproteinases (TIMPs)79. More recently, a comprehensive unbiased quantitative proteomic characterization of SASP led to the identification of additional and diverse SASP effectors, released as soluble molecules or in exosomes7981 with a set of components previously reported to be enriched in human plasma during ageing and age-associated diseases82. Exosomes were recently identified as key mediators of the paracrine senescence effects of SASP as well as of its protumorigenic properties83,84.

Exosomes

Extracellular vesicles produced by the endosomal compartment involved in intercellular communication.

Interplay between the DDR and SASP.

p16 induction can arrest the proliferation of normal cells and drive cellular senescence but is not sufficient to induce a complete SASP85. Of note, persistent DDR signalling is often required to initiate inflammatory cytokine secretion86. Consistent with a role for upstream DDR elements in promoting both cell-autonomous and paracrine functions of senescent cells, ATM, NBS1 and CHK2 prime SASP genes for activation, as depletion of these DDR upstream regulators dampens cytokine production in response to genotoxic stress86. Inhibition of p53 has the opposite effect, as it further enhances SASP following a senescence-inducing insult86, which may contribute to the generation of an inflammatory microenvironment that favours senescence escape and malignant transformation.

Recently, ATM was found to regulate expression of SASP genes indirectly by mediating the removal of the histone variant macroH2A1.1 from SASP genes in response to DNA damage and oncogenic stress87. However, because DDR activation is a quick response and SASP establishment is slow, additional pathways must control SASP. Indeed, the activation of the stress-inducible MAPK p38 was proven both necessary and sufficient to trigger growth arrest and SASP even in the absence of DNA damage88. Like ATM, p38 induces the expression of SASP transcripts by increasing the activity of nuclear factor-κB (NF-κB), suggesting that although DDR and p38 activation are independent, they can converge on SASP activation.

Ageing is one of the factors that is consistently associated with an increase in DDR in proliferating and non-proliferating cells, thus contributing significantly to the accumulation of senescent cells with age35,89. Incomplete DNA repair may further contribute to the accumulation of DNA lesions and DDR activation as well as to the widespread chromatin changes seen in different cell types and at the organismal level during ageing90,91. Moreover, DDR is a driver of metabolic reprogramming, which can enhance SASP92. Therefore, regulation of SASP by the DDR via multiple pathways can be one of the routes by which the DDR drives age-related inflammation.

Transcriptional and epigenetic control of SASP.

Several transcription factors and chromatin regulators have been implicated in the regulation of SASP, which is controlled mainly at the transcriptional level. NF-κB and the transcription factor CCAAT/enhancer-binding protein-β (C/EBPβ) bind promoters of SASP genes and regulate their activation74,93,94. Moreover, the transcription factor GATA4 activates a plethora of genes involved in immune response and inflammation, including those encoding IL-6, IL-8, CXCL1 (also known as GROα), granulocyte–macrophage colony-stimulating factor, and ECM proteases and their inhibitors. Because GATA4 levels increase following DDR activation and in senescent cells, it was proposed that GATA4 acts as the molecular link between DDR signalling and the subsequent NF-κB activation for full SASP establishment95,96, although GATA4 regulation of the NF-κB pathway happens indirectly through increased expression and secretion of IL-1α, an upstream regulator of NF-κB97.

The JAK2–STAT3 pathway activates a subset of SASP factors with immunosuppressive properties in a model of senescence induced by PTEN loss. In this context, JAK/STAT inhibitors were effective in reprogramming the SASP to enhance chemotherapy and T cell-mediated clearance of cancer senescent cells98. Furthermore, JAK inhibitors alleviated frailty in aged mice99. SASP gene expression is temporally dynamic100, and temporal changes in NOTCH1 activity during senescence have been reported to modulate the composition of SASP. NOTCH1 levels increase early on, and NOTCH1 activates transforming growth factor-β and its effectors, while keeping under control the proinflammatory arm of the SASP cascade by repressing its positive regulator C/EBPβ. At later stages, in deeply senescent cells, NOTCH1 levels are lower, and the SASP proinflammatory cytokines IL-1, IL-6 and IL-8 are induced94. Whether this function of NOTCH1 is linked to its recently reported role in direct ATM inhibition101,102 remains unknown.

The epigenetic reader bromodomain-containing protein 4 (BRD4), an acetylated histone-binding protein involved in oncogenesis, is recruited to superenhancers adjacent to SASP genes in OIS103. BRD4 contributes to the proper execution of cellular senescence and acts, unexpectedly, as a tumour suppressor. Indeed, chemical and genetic inactivation of BRD4 blunts the SASP, limits OIS immune-mediated cell clearance and may thus fail to remove damaged cells that are prone to senescence escape71. However, a recent chemical screen identified a small molecule that degrades BRD4 and has senolytic activity104. Moreover, BRD4 was shown to positively regulate telomere elongation in murine and human cultures105. Thus, BRD4 inhibitors may limit SASP activation but also cause more cellular senescence by promoting telomeric shortening. BRD4 binds to acetylated histone H3 Lys27, thus competing with Polycomb repressor complex 2 (PRC2), which methylates the same histone residue (producing trimethylated H3Lys27). Consistent with BRD4 and EZH2 (the catalytic core subunit of PRC2) competing for the same residue and having antagonistic functions, overexpression of EZH2 prevents entry into OIS through a variety of mechanisms, including DDR regulation and inhibition of the SASP gene expression programme106. The transcription-associated histone methyltransferase and oncoprotein MLL1 was also reported to be essential for SASP activation, but mainly through oncogene-induced hyper-replication and DDR engagement rather than direct transcriptional control of SASP genes107.

HMGB proteins also regulate the SASP. HMGB2 directly binds and specifically regulates SASP gene expression in oncogene-induced senescent cells, and its depletion diminishes SASP without affecting the senescence growth arrest108. HMGB1 functions mainly as one of the damage-associated molecular patterns, also known as alarmins109, that are released extracellularly to induce SASP-mediated paracrine senescence and alert the immune system110. Furthermore, increased nuclear pore density during OIS, which is key to establish SAHF, regulates SASP expression by mediating heterochromatin reorganization111.

HMGB proteins

Non-histone molecules that bind DNA and affect chromatin compaction.

SASP and innate immunity.

The DNA sensor cGAS and the adaptor protein STING have been reported to be major regulators of the SASP programme across species, presumably by activating NF-κB and the interferon response factor IRF3 (REFS60,112,113). cGAS–STING activation occurs mainly through recognition of ‘self’ double-stranded DNA or chromatin fragments in the cytosol of senescent cells114. cGAS–STING genetic depletion reduces the proinflammatory SASP and mitigates senescence immunosurveillance in senescence models in vivo. Adding to the complexity of SASP regulation, it was reported that aberrant activation of the cGAS–STING pathway could be linked to the downregulation of DNases (for example, DNase 2 and TREX1) that are normally devoted to cytoplasmic DNA degradation, resulting in cytosolic accumulation of nuclear DNA during senescence establishment115. While the mechanisms causing the release of cytosolic chromatin in senescent cells are still under intense scientific investigation, these discoveries suggest that STING inhibitors could potentially be used for the treatment of age-related chronic inflammation116.

cGAS–STING is not the only innate immunity pathway involved in SASP initiation and execution. The inflammasome, a multiprotein complex comprising caspase 1 and key regulators of the defence mechanisms against pathogens117, and Toll-like receptors, which regulate the inflammasome, have been reported to promote maturation and secretion of SASP factors during OIS73,118.

Senescent cells, especially when persisting for extended periods in culture or in vivo, display a profound activation of type I interferon response and downstream targets. This distinctively strong induction of type I interferon was found to be partly caused by reactivation of transposable elements and consequent to cGAS–STING activation62. Treatment with nucleoside reverse-transcriptase inhibitors, which inhibit retrotransposition of transposable elements, limits the senescence-associated detrimental effects of SASP and ameliorates chronic inflammation in aged animals62,119.

 

Cellular senescence of stem cells

Stem and progenitor cells are key to maintain tissue homeostasis and organization during physiological turnover and following tissue and organ injury. The functional capacity of stem cells, but not necessarily their number, declines with age120. DNA damage and markers of DDR activation have been observed in various stem cell types in different tissues and species, during normal and pathological ageing121,122, suggesting that stem cells are not immune to DNA damage accumulation and DDR activation123.

Although the activation of DDR pathways is expected to preserve genome stability and stemness, there is evidence that, following DNA damage, events regulated by the DDR lead to permanent cell cycle arrest with features of cellular senescence and cell differentiation. Indeed, exposure of mice to ionizing radiation leads to fur greying, which is due to damage-induced differentiation of hair bulb melanocyte stem cells after one round of cell division124. Although some markers of cellular senescence, such as p16 and SA-β-gal activity, were not detected, differentiation of melanocyte stem cells was associated with persistent DDR activation and was enhanced in Atm-knockout mice124. Similarly, self-renewal of haematopoietic stem cells (HSCs) is inhibited by telomere dysfunction or exogenous DNA damage followed by the induction of lymphoid differentiation125. Single-cell transcriptomic analysis of human HSCs revealed a dose-dependent activation of senescence-like programmes in response to DSBs, which were characterized by activation of p53 and the induction of proinflammatory programmes that resulted in reduced clonogenic potential, engraftment capacity and lineage output on transplantation126. Similarly, accumulation of DNA replication stress in HSCs from aged mice was linked to limited yet detectable HSC senescence, unbalanced haematopoietic differentiation and myeloid skewing127. However, especially in humans, it remains to be clarified whether myeloid-restricted haematopoiesis in elderly people results from increased myeloid-primed HSC differentiation or impaired lymphoid differentiation.

Myeloid skewing

An age-related proportional increase in myeloid cells at the expense of other lineages as observed in the bone marrow and blood.

DNA damage, for example induced by ionizing radiation, has been shown to promote differentiation and induce cellular senescence in mouse neural stem cells128. DNA damage led to a loss of expression of stemness genes and a general induction of a transcriptional profile indicative of differentiation into astrocytes. This cell differentiation programme was controlled by ATM and by soluble factors, in particular through BMP2 signalling128. Lineage tracing experiments in vivo in irradiated mice confirmed the induction of expression of differentiation markers in the subventricular zone of the brain, normally populated by neural stem cells128. Thus, DNA damage-induced cellular senescence can coincide with cell differentiation128,129.

Notably, in mouse embryonic stem cells, a p53-induced programme is associated with transcriptional activation of a differentiation programme and the repression of pluripotent stem cell genes130. Consistent with this observation, it was reported that senescence occurs in a programmed manner and contributes to mammalian embryonic development and tissue patterning, although mainly via induction of p21, p15 and mediators of the SASP rather than through overt DDR signalling131,132.

Altogether, these independent observations in melanocyte stem cells, HSCs, neural stem cells, embryonic stem cells and whole embryos suggest that persistent genotoxic stress in stem cells, and possibly more broadly in the context of progenitors and less differentiated cells, can lead to cellular senescence with features of cell differentiation. Although cellular senescence is not commonly considered a form of cell differentiation, as it is often the outcome of macromolecular damage, whereas cell differentiation is not, the commonalities are striking: they both involve cell cycle exit with a distinct transcriptional programme often controlled by soluble factors. It is possible that the common use, from the very beginning, of fairly differentiated cells (most typically fibroblasts) for senescence studies may have prevented the discovery of cellular senescence as a stress-induced differentiation programme, and perhaps with a different research history, cellular senescence could be known as a form of DNA damage-induced cell differentiation.

 

Can postmitotic cells become ‘senescent’?

The observation that during ageing terminally differentiated cells can accumulate persistent DNA damage and DDR markers35,89 prompts the question of whether persistent DNA damage signalling leads to the expression of cell cycle inhibitors and ultimately the establishment of cellular senescence, thus shifting cells from a non-dividing physiological state to a non-dividing pathological state. Although not extensively investigated, and mainly studied in neurons, evidence is emerging that cellular senescence can be associated with terminally differentiated cells133,134. Markers of DDR signalling, heterochromatin induction and activation of SASP, including the secretion of IL-6 and accumulation of SA-β-gal, were all detected in different types of neurons in ageing mice, and a short period of dietary restriction prevented their accumulation134. Such phenotypes were exacerbated in telomerase-inactivated mice, while p21 loss reduced many of these markers. Conversely, obesity in mice was associated with the expression of senescence markers in neurons of specific parts of the brain135. In a mouse model of Alzheimer disease, transcriptomic analyses of tau-containing neurons indicated an expression profile consistent with cellular senescence136. Neurons in the retina, known as retinal ganglion cells, were found to accumulate markers of DDR activation, SA-β-gal and both p16 and p21, in retinopathies137.

tau

A protein found in neurons that is important for maintaining microtubule structure in axons. Mutants and hyperphosphorylated forms are found in a variety of neurodegenerative diseases, including Alzheimer disease.

Mature postmitotic adipocytes show strong SA-β-gal staining and SASP induction in a p53-dependent manner in mice subjected to excessive caloric intake, although other markers, such as accumulation of DNA damage, were not reported138. Osteocytes are postmitotic differentiated cells controlling bone homeostasis. In old mice, osteocytes show markers of telomeric dysfunction and accumulate p16. SASP activation in osteocytes was shown to promote osteoclast activity, thus impairing bone strength. In this context, senolytic drugs reduced bone loss139.

Furthermore, persistent telomeric DNA damage in cardiomyocytes drives a senescence phenotype characterized by induction of p16 and p21 and a non-canonical SASP programme that contributes to cardiac hypertrophy and fibrosis140. Indeed, genetic and pharmacological clearance of p16-expressing senescent cells ameliorated heart functions in aged mice140. The most informative approach to determine the contribution of senescent cells to a condition is their genetic or pharmacological removal. However, as there are currently no tools to selectively target this subset of differentiated senescent cells, their role in ageing-related processes remains unclear.

Fibrosis

Pathological accumulation of extracellular matrix in diseases tissue that limits normal tissue function and leads to long-term tissue scaring.

Go to:

Beneficial effects of cellular senescence

Senescence can be regarded as a stress response that evolved to perform essential and beneficial functions141 (FIG. 3). The beneficial roles of senescent cells are evident in embryonic development. A distinct form of cellular senescence occurs in the mammalian developing embryo and in the placenta to control growth and patterning131,132,142,143. Similarly, in amphibians, cellular senescence occurs at specific steps during development to shape body growth144. Therefore, the cellular senescence early in life is important for normal development and morphogenesis, and later in life it becomes important for tissue repair and inhibition of cancer outgrowth. While tumour suppression activity is mediated mainly by cell-autonomous cell cycle arrest, most other senescence functions involve the SASP. Over time, although SASP favours proper tissue development, tissue repair and recruitment of immune cells, its persistence may generate chronic inflammation and contribute to ageing-related diseases and, paradoxically, cancer.

 

 

Fig. 3 |

Biological consequences of cell senescence.

Senescent cells execute distinct biological functions, which can have deleterious or beneficial consequences in a context-dependent manner. As beneficial functions, senescent cells guide tissue regeneration and embryonic development in the embryo in transient structures by secretion of FGF4 and FGF8 and shape the placenta structure and function with matrix metalloproteinase 2 and 9 (MMP2 and MMP9). Senescent cells also limit tissue damage by limiting excessive proliferation of cells and promote wound healing in part by secretion of PDGF-AA. One of the most prominent functions of senescence is tumour suppression. Senescent cells limit tumour development by cell-autonomous block of cell cycle progression via upregulation of p53, p16 and p21 and in a cell-non-autonomous manner by promoting senescence in neighbouring cells through secretion of interleukin-6 (IL-6) and IL-8. As deleterious functions, senescent cells can promote a proinflammatory microenvironment and therefore support tumour development in their proximity through multiple senescence-associated secretory phenotype (SASP) components. Similarly, senescent cells promote sterile chronic inflammation during ageing and during multiple age-related diseases. SASP factors, including IL-6, IL-1 receptor antagonist (IL-1RA), GROα and interferon-γ (IFNγ), are the main mediators of this effect. Additional SASP factors, including MMPs, might further damage tissue architecture and promote inflammation and tumorigenesis. When stem or progenitor cells enter senescence due to upregulation of the cell cycle inhibitory proteins, such as p16 and p21, they can no longer perform their function in supporting tissues by providing new cells, thus limiting tissue regenerative potential. Senescent cells also promote reprograming to an embryonic state, at least partially through IL-6. The reprograming, on one hand, can support tissue regeneration and, on the other hand, favours tumour development.

Cellular senescence contributes to the maintenance of the structure and the function of tissues following injury. For example, in liver fibrosis, a condition associated with scarring of the liver and decrease in its function, senescence limits the proliferation and expansion of ECM-producing activated hepatic stellate cells145,146. This response limits the progression of the pathology following liver damage. SASP from these cells attracts natural killer (NK) cells that eliminate senescent cells from the liver to restore liver homeostasis145. Cellular communication network factor 1 (CCN1; also known as CYR61), an ECM protein that mediates the induction of cellular senescence in the liver, promotes senescence in fibroblasts during cutaneous wound healing, thus limiting skin fibrosis147. During wound healing, the SASP component PDGF-AA accelerates wound closure12. Moreover, induction of a senescence programme limits fibrosis progression in the pancreas148.

Cellular senescence is involved in tissue repair in other systems. In zebrafish, senescence impairment prevents fin regeneration following amputation149, and in salamanders, senescent cells have been associated with limb regeneration150. Altogether, these observations suggest that cellular senescence is a programme that has evolved to limit tissue damage response in the organism and facilitate tissue repair and remodelling to promptly restore the tissue to a functional state.

The homeostatic function of senescent cells is dependent on their elimination by the immune system, once their beneficial functions have been performed145,151. Specialized SASP chemokines are able to attract distinct subsets of immune cells, including NK cells, neutrophils, dendritic cells, monocytes/macrophages, B cells and T cells75,151,152. Among these cell types, NK cells, T cells and macrophages can physically interact with senescent cells in pathological and physiological conditions131,132,145,150,153156. This response is mediated by SASP components and by the direct interaction between immune cells and senescent cells153,156158. The surveillance and clearance of senescent cells by the immune system is necessary in order to limit tumorigenesis in premalignant lesions and following cancer therapy155,156,159. Senescence immunosurveillance is also essential to limit pathological fibrotic conditions and ageing20,145,151,157. By contrast, during cancer development associated with accumulation of senescent cells, SASP can recruit immature myeloid cells to promote tumorigenesis in a paracrine manner160,161. In addition, through SASP, senescent cells contribute to cancer development162,163 and metastasis164, treatment failure and increased risk of recurrence165. Thus, SASP is a component of the senescence phenotype that seems to have evolved to signal the presence of senescent cells to the immune system and promote their elimination, but when senescent cells persist, their SASP becomes detrimental.

 

Detrimental impacts of senescence

Senescent cells can contribute to organismal ageing through multiple mechanisms (FIG. 3). With advancing age, individuals tend to develop a proinflammatory condition, characterized by high circulating levels of inflammatory molecules, known as inflammageing166. Inflammageing is a risk factor for various chronic age-associated diseases, including cardiovascular diseases, some cancer types and neurodegeneration, and can be associated with premature death. Furthermore, the presence of inflammatory molecules in the blood of elderly individuals is associated with weight loss, muscle loss and weakness, chronic inflammation and depression — manifestations of a condition recently recognized as frailty. A molecular link between cellular senescence, inflammageing and frailty was unveiled during the identification of common genomic variations that contribute to chronic phenotypes associated with ageing. Genome-wide association studies have recently revealed that the INK4/ARF gene locus, encoding p15INK4B, p16 and ARF, key effectors of the senescence growth arrest, is a genomic hotspot for susceptibility to several ageing-associated diseases, including cancer, diabetes and cardiovascular disease167,168, and physical dysfunction in people with advancing age169,170. Moreover, shortened telomere length, observed during ageing, correlates with metabolic and cardiac dysfunctions171,172. The SASP may contribute to dysfunction of multiple aged organs. Indeed, increased blood levels of IL-6, IL-1 receptor antagonist (IL-1RA) and tumour necrosis factor (TNF) receptor, all key SASP effectors, can be predictors of chronic disease in old individuals173. That cellular senescence is causative for frailty and age-associated diseases was demonstrated by transplantation of relatively few senescent cells and observation of tissue dysfunction and shortened lifespan in mice174.

Recently, atherosclerotic plaques from LDL receptor-null mice were found to accumulate high amounts of SA-β-Gal and p16-positive endothelial cells, vascular smooth muscle cells and macrophages. Removal of p16-positive cells in both p16-3MR and INK-ATTAC transgenic mice by both genetic methods and senolytics reduced plaque formation and progression, while dampening the SASP175. Consistent with these data in mice, human atherosclerotic plaques are highly enriched in p16-positive cells, although p16 seems to be expressed mainly by inflammatory macrophages176, and it cannot be ruled out that the positive effects of senolysis in atherosclerosis-prone mice175 are the consequence of elimination of inflammatory macrophages. Macrophages with senescence-like features have also been identified in lesions from patients with cell histiocytoses, haematological neoplasms associated with oncogene activation characterized by multiorgan dissemination of highly inflamed, p16-positive myeloid cells177,178.

LDL receptor

Mediates entry of LDL into cells. Mutations in the gene encoding this receptor predispose to the development of atherosclerosis.

In addition to the contribution of SASP to inflammation and chronic diseases at the organismal level, senescent cells might affect tissue regeneration by limiting the proliferative potential of stem and progenitor cells, as discussed earlier. Muscle progenitor cells that accumulate damage, both DNA and molecular, and upregulate p16 enter senescence on stimulation and are unable to contribute to muscle regeneration after injury179. From the findings taken together, senescence limits proliferation of stem and progenitor cells in a cell-autonomous manner. In addition, it was recently reported that the clonogenic properties of HSCs are impaired when they are exposed to SASP factors derived from senescent stromal cells180, suggesting that senescence may also affect regeneration in a paracrine fashion.

While the detrimental effects of senescence on adult somatic stem cell functions are well established, cellular senescence is more ambiguous in the context of somatic cell reprograming to an embryonic-like state. In vitro studies demonstrated that cellular senescence is a potent cell-autonomous barrier for transcription factor-mediated reprograming using OCT4, SOX2, KLF4 and MYC (commonly referred to as OSKM), in a manner similar to its role in tumour suppression181183. However, expression of these factors in vivo induces senescence and SASP production, which promotes paracrine senescence as well as reprogramming in non-senescent cells in a cell-non-autonomous manner184. In these same models, exogenous tissue damage that drives cellular senescence facilitates reprogramming. Senescence induction is necessary for efficient cell reprograming, as SASP factor production promotes reprogramming into induced pluripotent stem cells in a paracrine manner184. IL-6 produced in the SASP appears to be crucial for the generation of induced pluripotent stem cells in this context.

The non-cell-autonomous effects of cellular senescence can differ depending on the context. In vitro, the SASP factors IL-8, GROα, IL-6, and IGFBP7 reinforce the senescent phenotype in an autocrine manner74,93,185. In addition, specific SASP components, either soluble or in extracellular vesicles186, can induce senescence in a paracrine manner — which includes DDR activation73,83. These effects could potentially contribute to the spread of senescence in tissues and lead to tissue and organismal dysfunction. In vivo, SASP consequences are more complicated. For example, the SASP component interferon-γ induces cellular senescence and tissue ageing in mice with shortened telomeres, and abrogation of interferon-γ signalling rescues ageing-related phenotypes and extends lifespan187. Similarly, transforming growth factor-βimproves liver regeneration on injury by blocking paracrine senescence in neighbouring uninjured hepatocytes188. By contrast, short-term exposure to the SASP promoted expression of stem cell markers and increased the regenerative capacity of mouse keratinocytes189 and of skeletal muscle190 in vivo. However, prolonged exposure to the SASP resulted in paracrine-induced senescence, indicating that, at least in vivo, the effects are dependent on the composition and length of exposure to the SASP. From an evolutionary perspective, SASP may in the short-term facilitate wound healing and tissue damage repair by enhancing stem cell function, whereas the long-term presence of senescent cells may not be selected for and contributes to SASP deleterious effects.

 

Senescence as a driver of ageing

The rapidly ageing BubR1 hypomorphic mouse model has been valuable to demonstrate that cellular senescent cells cause ageing and disease191,192. BubR1 is part of the mitotic checkpoint machinery that ensures proper segregation of duplicated chromosomes into two identical daughter cells during mitosis. Mice expressing ~10% of normal BubR1 levels develop a variety of progeroid features, including shortened lifespan, cataracts, lordokyphosis, lipodystrophy and infertility very early in life191. Adipose tissue, skeletal muscle and the eyes of BubR1 hypomorphic mice express high levels of p16 and other senescence-associated features192. In an attempt to prevent the accumulation of these cells, BubR1-mutant mice were bred to Cdkn2ap16-knockout mice. In the absence of p16, the age-related deterioration of adipose tissue, skeletal muscle and eye was attenuated193. Importantly, genetically preventing the accumulation of p19ARF, a tumour suppressor that modulates the stability of p53 by influencing MDM2-mediated destruction, did not result in similar prevention193, indicating that p16 was critical for these disorders.

Cataracts

Clouding of the lens in the eye leading to inability to have clear vision. surgical intervention to replace diseased lenses is a common medical procedure in aged humans.

Lordokyphosis

Abnormal rearward curvature of the spine, observed both in laboratory mice and in humans.

Lipodystrophy

Abnormal distribution of adipose tissue in the body, can refer to both excessive or insufficient deposition.

On the basis of these observations, two different transgenic mouse models, INK-ATTAC11 and p16-3MR12, have been generated to critically test whether removal of senescent cells impacts ageing and diseases associated with senescent cell accumulation. Importantly, treatment of BubR1 hypomorphic mice harbouring the INK-ATTAC transgene to remove p16-expressing cells beginning at weaning age attenuated the accumulation of senescent cells and premature ageing in skeletal muscle, eye and adipose tissue11. A second study using the INK-ATTAC system in naturally aged mice corroborated these findings194 and increased median lifespan of both male and female mice in different genetic backgrounds and increased healthspan, as indicated by reduced kidney scarring, cardiomyocyte hypertrophy, cardiac stress intolerance, cataractogenesis and lipodystrophy194.

The p16-3MR mouse model expresses a trimodality reporter fusion protein consisting of synthetic Renilla luciferase, monomeric red fluorescent protein and a truncated herpes simplex virus thymidine kinase under the control of an artificial promoter for p16 (REF.12). In this model, cells that express p16 become sensitive to elimination by ganciclovir, a nucleoside analogue that is converted into a toxic DNA chain terminator by herpes simplex virus thymidine kinase and causes cell death195. These two mouse models have greatly accelerated our understanding of whether senescent cells contribute to ageing and age-related diseases, at least in model organisms, for numerous diseases, including Parkinson disease196, Alzheimer disease197,198, atherosclerosis175, idiopathic pulmonary fibrosis199, chronic obstructive pulmonary disease200 and osteoarthritis201. It remains unclear whether the elimination of senescent cells themselves or their SASP is the key element underlying these improvements.

Exploiting senescence for therapeutics

The literature is becoming inundated with evidence that senescent cells accumulate in a variety of age-associated diseases13. With the observation that elimination of senescent cells is largely beneficial and seems to lack long-term negative consequences, researchers in academia and industry have aimed to identify novel agents and strategies to eliminate senescent cells or their effects in the absence of genetical engineering to be applicable for use in humans. These ‘senotherapeutic’ strategies can be broadly categorized into two categories: pharmacological agents termed ‘senolytics’, which eliminate senescent cells, and senomorphics, which prevent the detrimental cell-extrinsic effects of senescent cells and include SASP inhibitors.

Senolytics.

Various senolysis strategies have been developed recently using a combination of in vitro models of senescence and in vivo animal models (FIG. 4TABLE 1). Senescent cells frequently upregulate negative modulators of apoptosis, including members of the BCL-2 family (including BCL-2, BCL-W and BCL-XL), which confers resistance to apoptosis-inducing signals14,15. The senolytic agents ABT-737 and ABT-263 (also known as navitoclax) inhibit the activity of the BCL-2 family members, thereby permitting senescent cells to initiate apoptosis. Additionally, A-1331852 and A-1155463, which are thought to inhibit BCL-XL, have also been shown to exhibit senolytic activity202. Recently, the cardiac glycoside ouabain demonstrated senolytic activity, at least in part, through the induction of NOXA, a proapoptotic BCL-2 family protein203,204. Promoting proteasomal degradation of BCL-2 through EF24 treatment also results in the selective killing of senescent cells205. Administration of proxofim (a peptide) promotes senolysis by interfering with the binding of p53 to forkhead box protein O4 (FOXO4)206. In senescent cells, FOXO4 binds to p53 to localize it to the nucleus. If this interaction is disrupted by the administration of an inverse peptide, p53 is excluded from the nucleus, initiating cytochrome c release from the mitochondria and apoptosis206. The use of various natural flavonoids, including quercetin and fisetin, either alone or in combination with dasatinib, a pan-tyrosine kinase inhibitor, can stimulate senolysis in a variety of contexts in vitro and in vivo199,207210. In agreement with studies performed on INK-ATTAC naturally aged mice194, treatment of advanced-age mice with a combination of dasatinib and quercetin improved physical function and increased lifespan174. Importantly, the administration of dasatinib and quercetin has shown effectiveness in reducing the expression of p16 and SA-β-gal in a phase I clinical trial among patients with diabetic kidney disease211 and idiopathic pulmonary disease212. Other senolytics, including HSP90 inhibitors213,214 and piperlongumine215,216, have also been demonstrated to be selective towards senescent cells. More recently, clinically approved antibiotics have been reported to have senolytic activity towards DNA-damage induced senescent cells through metabolic changes217. Collectively, these strategies target a broad spectrum of cellular pathways, indicating that senescent cells can be removed via multiple avenues.

 

 

Fig. 4 |

Senolytic therapeutic interventions.

The sensitivities of senescent cells to pharmacological treatments that can promote their death are diverse. A number of known mechanisms of senolytic action are indicated; the various specific compounds that hit these nodes are indicated. Impacting tyrosine kinase (TK) through the use of dasatinib (when used either alone or in combination with the flavonoid quercetin) is capable of initiating death of certain senescent cell types. Quercetin and fisetin are natural flavonoids that impact mammalian target of rapamycin (mTOR) signalling. Inhibitors of the antiapoptotic members of the BCL-2 family are capable of inducing death through mitochondrial-mediated mechanisms, which can also be elicited by the action of cardiac glycosides such as ouabain. Inhibitors of HSP90 or histone deacetylases (HDAC) have also been suggested to promote selective apoptosis of senescent cells. Additionally, disruption of binding of forkhead box protein O4 (FOXO4) to p53, which occurs in senescent cells, through the use of a small peptide liberates p53 to activate apoptosis. Galactose-conjugated senolytic prodrugs (GAL-prodrugs) are processed by senescence-associated-β-galactosidase (SA-β-gal) to exert selective senescence targeting. RTK, receptor tyrosine kinase.

Table 1 |

Senolytic and senomorphic compounds

Compound

Targets

Clinical trial status

Refs

Senolytic

     

Dasatinib

Pan-receptor tyrosine kinases (including ephrin B1)

Phase II, (NCT02848131) for chronic kidney disease, phase II (NCT04313634) for skeletal health, phase I/II (NCT04063124) for Alzheimer disease

210

Quercetin

Numerous (including PI3K)

Phase II NCT02848131) for chronic kidney disease, phase II (NCT04313634) for skeletal health, phase I/II (NCT04063124) for Alzheimer disease

210

Fisetin

PI3K/AKT/mTOR

Pphase I/II (NCT04210986) for osteoarthritis of the knee, phase II (NCT04313634) for skeletal health

202

ABT-737

BCL-2, BCL-XL and BCL-W (prosurvival proteins)

Preclinical animal models

14

ABT-263 (navitoclax)

BCL-2, BCL-XL and BCL-W (prosurvival proteins)

Phase I/II (NCT00445198), phase II (NCT02591095), phase I (NCT02520778), phase II (NCT02079740) for various cancers

15,165,175,197,228

A-1331852

BCL-XL (prosurvival protein)

Preclinical models of senescence in vitro

202

A-1155463

BCL-XL (prosurvival protein)

Preclinical models of senescence in vitro

202

EF24

Proteosomal degradation of BCL-2 family proteins (prosurvival proteins)

Preclinical models of senescence in vitro

205

Cardiac glycosides (including oubain and digoxin)

BCL-2, BCL-XL and BCL-W (prosurvival proteins)

Preclinical animal models

203,204

Azithromycin

Autophagy, metabolic changes

Preclinical models of senescence in vitro

217

Roxithromycin

Autophagy, metabolic changes

Preclinical models of senescence in vitro

217

Proxifim

p53

Preclinical animal models

206

UBX0101

MDM2 and p32

Phase II (NCT04129944) for osteoarthritis of the knee

201

Panobinostat

HDAC

Approved for multiple myeloma

55

Geldanamycin

HSP90

Preclinical models of senescence in vitro

214

Tanespimycin

HSP90

Preclinical models of senescence in vitro

214

Alevspimycin (17-DMAG)

HSP90

Preclinical models of senescence in vitro

214

Piperlongumine (and analogues)

OXR1 (unknown)

Preclinical models of senescence in vitro

215,216

Galactose-conjugated nanoparticles

Lysosomal activity of senescent cells

Preclinical animal models

218

Galactose-modified cytotoxic agents

Lysosomal activity of senescent cells

Preclinical animal models

219,220

BET protein degrader

Bromodomain and extraterminal domain family protein

Preclinical animal models

104

Senomorphic

     

Metformin

IKK and/or NF-κB

Approved for type 2 diabetes

137,235237

Apigenin

NF-κB p65 subunit and IκB

Naturally occurring flavonoid

238

Kaempferol

NF-κB p65 subunit and IκB

Naturally occurring flavonoid

238

BAY 11–7082

NF-κB p65 subunit and IκB

Preclinical models of senescence in vitro

118

Rapamycin

mTOR

Approved for immunosuppression

230,231,233,234

RAD001

mTOR

Approved for tuberous sclerosis complex-associated diseases

232

SB203580

p38 MAPK

Preclinical models of senescence in vitro

88

(5Z)-7-Oxozeaenol

TAK1

Preclinical models of senescence in vitro

232

Ruxolitinib

JAK

Approved for graft-versus-host disease

99

Senomorphic (cont.)

     

KU-60019

ATM

Preclinical animal models

274

NDGA

HSP90

Naturally occurring antioxidant

249

Loperamide

HSP90

Approved for treatment of diarrhoea

214

Simvastatin

IL-6, IL-8, MCP1

Preclinical models of senescence in vitro

240

Cortisol

IL-6 secretion

Steroid hormone

241

Anakinra

IL-1R

Approved for rheumatoid arthritis

242

Canakinumab

IL-1β

Approved for cryopyrin-associated periodic syndromes

243

Rilonacept

IL-1α and IL-1β

Approved for cryopyrin-associated periodic syndromes

244

Etanercept

TNF

Approved for autoimmune diseases

245

Infliximab

TNF

Approved for autoimmune diseases

246

Tocilizumab

IL-6R

Approved for autoimmune diseases

247

Siltuximab

IL-6

Approved for multicentric Castleman disease

248

Telomeric antisense oligonucleotides

Telomeric non-coding RNA fuelling DDR

Preclinical animal models

251

Open in a separate window

ATM, ataxia telangiectasia mutated; DDR, DNA damage response; HDAC histone deacetylase; IL, interleukin; IL-1R, interleukin-1 receptor; IL-6R, interleukin-6 receptor; IκB, inhibitor of nuclear factor-κB; IKK, inhibitor of nuclear factor-κB kinase; JAK, Janus kinase; MCP1, monocyte chemoattractant protein 1; mTOR, mammalian target of rapamycin; nuclear factor-κB; NF-κB; NGDA, nordihydroguaiaretic acid; TNF, tumour necrosis factor.

HSP90

A molecular chaperone that promotes proper protein folding and degradation, which also contributes to heat stress resilience.

A novel strategy to induce senolysis exploits the observed increased level of SA-β-gal activity. Nanoparticles containing either fluorophores or cytotoxic agents coated with galacto-oligosaccharides were found to preferentially deliver cytotoxic cargo to senescent cells because of the higher level of SA-β-gal activity in these cells218. Additionally, recent studies have further shown the potential of delivering cytotoxic factors to lysosomes of senescent cells by galactose-modifying prodrugs or cytotoxic agents219,220.

Prodrugs

Compounds that are metabolized into an active drug to modify drug bioavailability and activity.

In support of senolytic therapies being potentially beneficial is the notion that organisms have an intrinsic senolytic system: immunosurveillance against senescent cells. Senescent cells are indeed subjected to immunosurveillance by multiple components of innate and adaptive immunity, including NK cells, T cells and macrophages145,151,153156. Therefore, it is conceivable to harness the mechanisms of immunosurveillance of senescent cells, and all mechanisms that the immune system uses to target other threats, to eliminate senescent cells. Several such approaches were recently implemented (reviewed in221). It is possible that senescent cells accumulate in aged and diseased tissues owing to a decline in immunosurveillance. Therefore, restoring or boosting the ability of the immune system to specifically eliminate senescent cells could result in their successful clearance from tissues. Such an approach is based on our understanding of the mechanisms of immunosurveillance of senescent cells, in particular the interaction between NK and senescent cells151,157,158. NK cells use perforin-mediated granule exocytosis and not death receptor ligands, which induce cell death by binding to these receptors, to target senescent cells151. This mechanism is favoured because decoy receptor 2 (DCR2) is strongly expressed in senescent cells. DCR2 prevents targeting via the death receptors (DR4 and DR5) by the variety of cytotoxic cells that express their ligand TRAIL151. Therefore, blocking such an inhibitory mechanism can lead to removal of the inhibitory effect and increased targeting of senescent cells by endogenous naturally occurring mechanisms.

Perforin

A pore forming protein expressed in cytotoxic T cells and natural killer cells. When these cells execute cytotoxicity, they secrete granules containing perforin, which binds to the target cell’s membrane and forms pores on the target cell in order to allow cytotoxicity.

An alternative approach to enhance the immune clearance of senescent cells is by enhancing the activity and increasing the accumulation of immune cells that are responsible for senescent cell surveillance. Stimulation of innate immune response with poly(I:C), a simulator of viral infection, improves senescent cell clearance145. Although treatment with similar agents can hardly be considered in humans due to possible side effects, the subtler stimulation of the immune system with specific cytokines that boost NK cells could be a plausible approach. The cytokines IL-21 and IL-15 have been suggested to substantially boost NK cell-mediated immunity against cancer cells222225. However, the effect of these cytokines on immunosurveillance of senescent cells has not been elucidated and their efficacy as senolytics in disease models needs to be tested.

While boosting natural immune mechanisms of senescent cell surveillance may have therapeutic potential in years to come, it is worth considering the rapidly growing variety of tools that are becoming available in the immuno-oncology field. Directed cellular approaches such as use of chimeric antigen receptor T (CAR T) cells and NK cells, as well as blocking of immunoinhibitory interactions by blocking interactions with PD1, cytotoxic T lymphocyte-associated protein 4 (CTLA4) and other inhibitory molecules might provide powerful strategies for increasing immunosurveillance. Such approaches depend on the recognition of specific markers on the cell surface of senescent cells. Exciting recent work has demonstrated that senolytic CAR T cell therapy can attenuate senescence-associated diseases226. Several studies used unbiased approaches to identify such markers227229. However, the overlap between the extracellular markers identified by the aforementioned strategies is low, when each experiment identified distinct markers, suggesting that such markers might be specific to the cell of origin and/or to the mechanism of senescence induction. One possible way to address this problem is to use the surface molecules on senescent cells that signal to NK cells for their elimination, namely the NKG2D receptor ligands157. However, the large repertoire of such ligands, including MICA, MICB and ULBP1–ULBP6, and the different levels of their expression on senescent cells of different origins157 might provide a challenge in using them for enhancing immune clearance of senescent cells. When different markers are present on cells of different origins, it is possible to envision multiple approaches specific to distinct pathological conditions.

Chimeric antigen receptor T (CAR T) cells

T cells that have been genetically engineered to express T cell receptor developed to bind a defined target in order to eliminate the cells that have the target on their membrane.

PD1

A protein expressed on the cell surface that inhibits the ability of the immune system to target the cells that express the protein. Inhibition of interaction of PD1 with its ligand is a potent immunotherapy approach.

Senomorphics.

An alternative to the complete elimination of senescent cells through senolysis is the use of senomorphic agents. The principle of senomorphics is to disrupt key attributes of senescence, primarily SASP production and secretion, while keeping the cells alive, or to modify their ability to maintain a stable growth arrest (FIG. 5). This approach could interfere with the proinflammatory nature of senescent cells and potentially delay key aspects of ageing and ageing-associated disease.

 

 

Fig. 5 |

Senomorphic therapeutic interventions.

As an alternative to active killing of senescent cells, senomorphic approaches try to limit the detrimental impacts of these cells, largely through modulation of the senescence-associated secretory phenotype (SASP). As for senolytics, a number of common nodes have been identified that may be unique opportunities for intervention. Rapamycin, a well-characterized inhibitor of mammalian target of rapamycin (mTOR), has been shown to increase lifespan of laboratory mice. Additionally, rapamycin decreases production of the SASP, which may explain the beneficial impacts on life. Nuclear factor-κB (NF-κB) is a critical component for SASP production, and inhibition of NF-κB activity decreases the ability of cells to be proinflammatory. Additionally, inhibition of HSP90 is able to modulate SASP production. Similarly, Janus kinase (JAK)/signal transducer and activator of transcription (STAT) inhibitors and blocking of interleukin-6 (IL-6), IL-1 and tumour necrosis factor (TNF). Taken together, these molecules are beginning to elucidate ways that proinflammatory signalling from senescent cells can be attenuated in the hope of decreasing the consequences of senescent cell accumulation in tissues. ASO, antisense oligonucleotides; DDR, DNA damage response; DDRNA, DNA damage response RNA; dilncRNA, damage-induced long non-coding RNA; NDGA, nordihydroguaiaretic acid; RTK, receptor tyrosine kinase.

Novel mechanisms to regulate the SASP, in addition to the transcriptional modulation of SASP factor expression, have been uncovered by using hypothesis-driven strategies and elegant genetic and drug screenings. These include, for example, the mammalian target of rapamycin (mTOR) pathway, which coordinates cell growth and metabolism in response to nutrients, and also promotes SASP production through increased translation of subsets of mRNAs, including the membrane-bound and upstream regulator of NF-κB, IL-1α230, and the serine/threonine kinase MK2, which indirectly stabilizes many cytokine-encoding transcripts231. These mechanistic insights provided a molecular foundation for the use of the mTOR inhibitor rapamycin in pathological settings associated with senescence in vivo. Treatment with rapamycin (and its analogue RAD001)232 attenuated the protumorigenic SASP230, prevented senescence233, impaired SASP-mediated immune recognition of oncogene-expressing cells and ameliorated liver dysfunction in naturally aged mice231. Of note, it cannot be excluded that rapamycin may act also via senescence-independent mechanisms. Additionally, rapamycin treatment increased lifespan and delayed certain ageing-related dysfunctions in mice234.

Compounds that modulate NF-κB signalling, including metformin137,235237, apigenin238, kaempferol238 and BAY 11–7082 (REF.118), have also been shown to decrease SASP production. NAD+/NADH metabolism was identified as a critical regulator of the magnitude of proinflammatory SASP associated with oncogene activation, and this regulation can be independent of senescence-induced growth arrest239. A number of neutralizing antibodies directed against key components of the SASP or their receptors, including IL-6, IL-1α, IL-1β and TNF, also have shown senomorphic properties104,240248 (TABLE 1). Additionally, SASP establishment and the secretion of SASP factors can be modulated by inhibiting HSP90 (REFS214,249).

Finally, as in many instances cellular senescence is the consequence of the activation of DDR pathways by dysfunctional telomeres, the inhibition of telomeric DDR may prevent or reduce senescence establishment and maintenance. Recently, sequence-specific inhibition of DDR activation by antisense oligonucleotides (ASOs)250 and their use in cultured cells and in mouse models to specifically inhibit telomeric DDR38 provided support for this approach. The use of telomeric ASOs in a mouse model of Hutchinson–Gilford progeria syndrome (an accelerated ageing syndrome) effectively reduced DDR activation, the levels of senescence markers and SASP induction, improved tissue homeostasis and extended lifespan251. This or similar approaches that do not depauperate stem or progenitor cell reservoirs but rather promote cell proliferation may provide an alternative or complementary approach to senolysis.

Senolytics may have benefits compared with senomorphics. Firstly, removal of senescent cells has the advantage of their being targeted intermittently and not requiring continuous administration of SASP inhibitors, although repeated treatments are probably necessary. Furthermore, the removal of senescent cells eliminates the possibility of senescence bypassing mutations that can promote tumorigenesis in these damaged cells. Additionally, although there is a strong correlation between the SASP and ageing-associated tissue and organ dysfunction, there is no direct demonstration that the SASP drives these ageing-related defects, as it has not been possible to separate the SASP from senolysis using transgenic mouse models. However, although INK-ATTAC transgenic mouse models in which senescent cells are removed exhibit no apparent detrimental side effects, it remains to be determined whether prolonged or repeated senolysis could become eventually toxic to an organism. Furthermore, it is not known whether senolysis is detrimental or beneficial when the senolysis is induced in advanced age, when the organism has a high senescent cell burden. Pharmacological clearance of a subpopulation of p16/SA-β-gal-positive macrophages has been proposed to contribute to the beneficial effects of senolysis252,253, but the extent to which this cell type contributes to age-related dysfunctions needs to be further investigated. Lastly, emerging evidence suggests that targeted senolysis in mice may be profoundly toxic in the liver and perivascular tissue, because of the eradication of p16-expressing endothelial cells, adipocytes and macrophages in aged organs, which all have structural functions254.

Interplay between ‘rejuvenating’ treatments and cellular senescence.

Caloric restriction has been demonstrated to be the most effective strategy to lengthen healthspan and lifespan, and is efficacious in a range of species from yeast to primates255. Whether it impacts the number or activity of senescent cells has been surprisingly understudied. Nevertheless, it was reported that caloric restriction reduces p16 levels and the transcriptional expression of genes associated with cellular senescence, including SASP genes, in mice and in the colon of healthy human individuals256,257. Caloric restriction in mice reduced the DDR and improved telomere maintenance258. Caloric restriction has also been found to reduce the levels of DDR markers and SASP regulators in postmitotic neurons134. The link between caloric restriction and reduced DDR signalling and reduced senescence burden may be related to the observation in culture that serum boosts DDR signalling in senescent cells259.

Inhibiting telomere shortening to prevent and reduce cellular senescence in the context of ageing and so-called telomere syndromes is being considered as a therapeutic approach260. In addition to potentially using telomeric ASOs to modulate DDR activation and its consequences251, other options are being explored, such as the reactivation of an endogenous telomerase gene using a natural compound261, but with limited efficacy, and by sex hormones262, which however has some significant clinical drawbacks. Viral delivery of the telomerase-encoding gene (Tert) has been tested with more success in several settings. Systemic delivery of Tert reduces several senescence markers and ageing-related conditions and extends the lifespan of wild-type mice, thus demonstrating that telomeric DDR activation plays a role in natural ageing263. Idiopathic pulmonary fibrosis is associated with telomere shortening and markers of cellular senescence in humans264. In mouse models recapitulating these features, adeno-associated virus particles delivering Tert have been shown to reduce DDR and the level of senescence markers and improve lung function265. Importantly, concerns regarding the safety of telomerase forced expression in damaged tissues, in particular by fuelling cancer progression, have been mitigated by the recent demonstration that oncogene-expressing mice do not show accelerated tumorigenesis on such treatment266.

Over the past few decades, heterochronic parabiosis, a process by which young and old small animals are surgically connected by establishing a joined circulation, identified systemic factors that are present in young blood and can ameliorate the function of several aged organs, including liver, muscle, heart and brain267. In the brain, similar rejuvenating effects were observed by a simpler procedure of young human plasma transferred into old animals268. Exposure to a younger systemic environment was also shown to relieve age-associated tissue dysfunctions in ageing telomerase-deficient mice269. Recently, it was reported that blood exchange between young and old mice led to a significant reduction in cellular senescence and SASP marker expression in multiple aged tissues, while at the same time the levels of senescence markers were increased in the young animals exposed to old blood270. These observations indicated that systemic factors reverse some features associated with ageing, including defective stem and progenitor cell function, chronic inflammation and senescence burden, and support the hypothesis that intermittent blood exchange in humans may be used as a therapeutic modality for age-related diseases. In line with this, therapeutic plasma exchange is currently being tested in patients with acute sepsis271 or liver damage272.

Challenges and future directions

As our understanding of senescent cell characteristics in vitro and in vivo continues to increase, many challenges remain. For example, it is unclear how many ‘senescent phenotypes’ exist. It is possible that there is a very high degree of heterogeneity in the senescent state, at the single-cell level, as well as between cell types and depending on the stimuli that induce senescence. The emerging notion that cellular senescence is a dynamic process that evolves over time further augments its complexity. The identification of a truly universal marker of senescence would be a boon to isolate and characterize senescent cells. The identification of more specific markers to distinguish unequivocally different types of senescent cells would be most useful to characterize them and shed light on their origin in vivo. Presently, single-cell transcriptomic approaches, including spatial transcriptomics, are the only option to fully appreciate senescent cell complexity and to determine the similarities and differences between regulated processes such as cell differentiation and senescence and the impact that senescence has on already differentiated, non-proliferating, cells. Distinguishing between senescent cell subtypes and identifying what triggers senescence for each sub-type would enable us to identify the particular subsets of senescent cells that may be most deleterious to tissue function, and their targeting would optimize the benefits of senolytic and senomorphic approaches, while minimizing deleterious effects.

The physiological triggers of senescence in vivo are poorly understood. Telomere dysfunction is probably an important one, and only the use of specific telomeric DDR inhibitors will prove its involvement in various physio-pathological conditions. Telomeric ASOs could be both a research tool and a potential treatment, selective for those forms of cellular senescence caused by telomere damage.

The triggers for senescence in seemingly DDR-independent conditions remain elusive. SASP is emerging as the most consequential of all senescent cell phenotypes. Yet, the appreciation of the complexity underlying the mechanisms that control the SASP evolves as more components of the SASP are identified, in different cell types and in different contexts. Also to be considered is the powerful paracrine impact of cellular senescence on non-senescent cells; controlling it will clarify its expected but undemonstrated role in organismal ageing and in a number of conditions associated with the spread of cellular senescence.

Much of our understanding of the contribution of senescent cells to disease comes from animal models for human conditions. However, it remains to be shown that senolysis in humans is safe or effective, which is a prerequisite for devising treatments for patients. The study of long-term effects in rodents is limited to 2–3 years after senolysis, which is a much shorter time frame than their potential use in humans. Therefore, possible long-term toxic effects or negative consequences of senescent cell elimination that require longer observation times simply cannot be assessed with our current models and tools. It is clear that the immune system has the capacity to eliminate senescent cells in certain contexts; however, the clearance process seems to become dysfunctional with age and in disease, possibly explaining the accumulation of senescent cells with age273. As the cell components of the immune system are also subjected to senescence, it will be important to determine whether senotherapies eliminate these cells and whether the elimination of senescent immune cells contributes to or mediates the effects of senotherapies. In addition, harnessing the intrinsic ability of the immune system to target these cells, perhaps through engineered T cells, may hold promise for novel therapeutics. In summary, cellular senescence is clearly more complex and nuanced than initially thought, making diverse and occasionally contrasting contributions to physiology and ageing. Importantly, the many years of basic research in this field have set the foundations for a now exploding biotech and industrial activity devoted to turning such knowledge into treatments for patients. The next few years will see whether its promise is fulfilled: exciting times lie ahead.

 

Acknowledgements

R.D.M. is supported by Telethon (TIGET grant E5), a Career Development Award from the Human Frontier Science Program, an Advanced Research Grant from the European Haematology Association, “Pilot and Seed Grant 2015” from San Raffaele Hospital, a Hollis Brownstein Research Grant from the Leukemia Research Foundation, the Interstellar Initiative on Healthy Longevity of the New York Academy of Sciences and the Japan Agency for Medical Research and Development, and the Italian AIRC under MFAG 2019, ID. 23321 project PI Di Micco Raffaella. V.K. is supported by grants from the European Research Council under the European Union’s Seventh Framework Proframme (309688) and under Horizon 2020 (856487), from the Israel Science Foundation (634-15; 2633-17), from the Israel Ministry of Health, Minerva Center “Ageing, from Physical Materials to Human Tissues”, and from the Sagol Institute for Longevity Research. V.K. is an incumbent of the Georg F. Duckwitz Professorial Chair. D.B. is supported by the Ellison Medical Foundation, the Glenn Foundation for Medical Research, the US National Institutes of Health (R01AG053229 and R01AG068076), the Mayo Clinic Children’s Research Center, the Alzheimer’s Disease Research Center at Mayo Clinic and the Cure Alzheimer’s Fund. F.d’A.d.F. is supported by AIRC (application 12971), Fondazione Telethon (GGP17111), PRIN 2015, the InterOmics Project, the AMANDA project Accordo Quadro Regione Lombardia-CNR, a European Research Council advanced grant (322726), a European Research Council proof-of-concept grant (875139), AriSLA (project “DDRNA and ALS”), the AIRC Special Program 5 per mille metastases (Project-21091) and the European Joint Programme on Rare Diseases (EJP RD).The authors apologize for not being able to cite all the important contributions of their colleagues owing to space limitations.

 

References

  1. Hayflick L & Moorhead PS The serial cultivation of human diploid cell strains. Exp. Cell. Res25, 585–621 (1961). [PubMed] [Google Scholar]
  2. Hayflick L The limited in vitro lifetime of human diploid cell strains. Exp. Cell. Res37, 614–636 (1965). [PubMed] [Google Scholar]
  3. Serrano M, Hannon GJ & Beach D A new regulatory motif in cell-cycle control causing specific inhibition of cyclin D/CDK4. Nature366, 704–707 (1993). [PubMed] [Google Scholar]
  4. Alcorta DA et al. Involvement of the cyclin-dependent kinase inhibitor p16 (INK4a) in replicative senescence of normal human fibroblasts. Proc. Natl Acad. Sci. USA93, 13742–13747 (1996). [PMC free article] [PubMed] [Google Scholar]
  5. Serrano M et al. Role of the INK4a locus in tumor suppression and cell mortality. Cell85, 27–37 (1996). [PubMed] [Google Scholar]
  6. Serrano M, Lin AW, McCurrach ME, Beach D & Lowe SW Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell88, 593–602 (1997). [PubMed] [Google Scholar]
  7. Di Micco R et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature444, 638–642 (2006). [PubMed] [Google Scholar]
  8. Bartkova J et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature444, 633–637 (2006). [PubMed] [Google Scholar]
  9. Campisi J Cellular senescence and apoptosis: how cellular responses might influence aging phenotypes. Exp. Gerontol38, 5–11 (2003). [PubMed] [Google Scholar]
  10. Giaimo S & d’Adda di Fagagna F Is cellular senescence an example of antagonistic pleiotropy? Aging Cell11, 378–383 (2012). [PubMed] [Google Scholar]
  11. Baker DJ et al. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature479, 232–236 (2011). [PMC free article] [PubMed] [Google Scholar]This study demonstrates that selective elimination of p16-expressing senescent cells is safe and capable of modulating age-related dysfunction in a premature aged mouse model.
  12. Demaria M et al. An essential role for senescent cells in optimal wound healing through secretion of PDGF-AA. Dev. Cell31, 722–733 (2014). [PMC free article] [PubMed] [Google Scholar]
  13. Childs BG et al. Senescent cells: an emerging target for diseases of ageing. Nat. Rev. Drug Discov16, 718–735 (2017). [PMC free article] [PubMed] [Google Scholar]
  14. Yosef R et al. Directed elimination of senescent cells by inhibition of BCL-W and BCL-XL. Nat. Commun7, 11190 (2016). [PMC free article] [PubMed] [Google Scholar]
  15. Chang J et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat. Med22, 78–83 (2016). [PMC free article] [PubMed] [Google Scholar]
  16. Childs BG, Baker DJ, Kirkland JL, Campisi J & van Deursen JM Senescence and apoptosis: dueling or complementary cell fates? EMBO Rep. 15, 1139–1153 (2014). [PMC free article] [PubMed] [Google Scholar]
  17. Kirschner K et al. Phenotype specific analyses reveal distinct regulatory mechanism for chronically activated p53. PLoS Genet. 11, e1005053 (2015). [PMC free article] [PubMed] [Google Scholar]
  18. Milanovic M et al. Senescence-associated reprogramming promotes cancer stemness. Nature553, 96–100 (2018). [PubMed] [Google Scholar]
  19. Herranz N & Gil J Mechanisms and functions of cellular senescence. J. Clin. Invest128, 1238–1246 (2018). [PMC free article] [PubMed] [Google Scholar]
  20. Ovadya Y et al. Impaired immune surveillance accelerates accumulation of senescent cells and aging. Nat. Commun9, 5435 (2018). [PMC free article] [PubMed] [Google Scholar]
  21. Jackson SP & Bartek J The DNA-damage response in human biology and disease. Nature461, 1071–1078 (2009). [PMC free article] [PubMed] [Google Scholar]
  22. Fumagalli M, Rossiello F, Mondello C & d’Adda di Fagagna F Stable cellular senescence is associated with persistent DDR activation. PLoS ONE9, e110969 (2014). [PMC free article] [PubMed] [Google Scholar]
  23. Galbiati A, Beausejour C & d’Adda di Fagagna F A novel single-cell method provides direct evidence of persistent DNA damage in senescent cells and aged mammalian tissues. Aging Cell16, 422–427 (2017). [PMC free article] [PubMed] [Google Scholar]
  24. d’Adda di Fagagna F et al. A DNA damage checkpoint response in telomere-initiated senescence. Nature426, 194–198 (2003). [PubMed] [Google Scholar]This study demonstrates that the activation of the DDR pathways by critically short telomeres is key in the enforcement of cellular senescence.
  25. Mallette FA & Ferbeyre G The DNA damage signaling pathway connects oncogenic stress to cellular senescence. Cell Cycle6, 1831–1836 (2007). [PubMed] [Google Scholar]
  26. Beausejour CM et al. Reversal of human cellular senescence: roles of the p53 and p16 pathways. EMBO J. 22, 4212–4222 (2003). [PMC free article] [PubMed] [Google Scholar]
  27. Dulic V, Beney GE, Frebourg G, Drullinger LF & Stein GH Uncoupling between phenotypic senescence and cell cycle arrest in aging p21-deficient fibroblasts. Mol. Cell. Biol20, 6741–6754 (2000). [PMC free article] [PubMed] [Google Scholar]
  28. Kamijo T et al. Tumor suppression at the mouse INK4a locus mediated by the alternative reading frame product p19ARF. Cell91, 649–659 (1997). [PubMed] [Google Scholar]
  29. Sherr CJ Divorcing ARF and p53: an unsettled case. Nat. Rev. Cancer6, 663–673 (2006). [PubMed] [Google Scholar]
  30. Kamijo T et al. Loss of the ARF tumor suppressor reverses premature replicative arrest but not radiation hypersensitivity arising from disabled atm function. Cancer Res. 59, 2464–2469 (1999). [PubMed] [Google Scholar]
  31. Velimezi G et al. Functional interplay between the DNA-damage-response kinase ATM and ARF tumour suppressor protein in human cancer. Nat. Cell Biol15, 967–977 (2013). [PubMed] [Google Scholar]
  32. Herbig U, Jobling WA, Chen BP, Chen DJ & Sedivy JM Telomere shortening triggers senescence of human cells through a pathway involving ATM, p53, and p21(CIP1), but not p16(INK4a). Mol. Cell14, 501–513 (2004). [PubMed] [Google Scholar]
  33. Hemann MT, Strong MA, Hao LY & Greider CW The shortest telomere, not average telomere length, is critical for cell viability and chromosome stability. Cell107, 67–77 (2001). [PubMed] [Google Scholar]
  34. Bodnar AG et al. Extension of life-span by introduction of telomerase into normal human cells. Science279, 349–352 (1998). [PubMed] [Google Scholar]
  35. Fumagalli M et al. Telomeric DNA damage is irreparable and causes persistent DNA-damage-response activation. Nat. Cell Biol14, 355–365 (2012). [PMC free article] [PubMed] [Google Scholar]
  36. Hewitt G et al. Telomeres are favoured targets of a persistent DNA damage response in ageing and stress-induced senescence. Nat. Commun3, 708 (2012). [PMC free article] [PubMed] [Google Scholar]
  37. Bae NS & Baumann P A RAP1/TRF2 complex inhibits nonhomologous end-joining at human telomeric DNA ends. Mol. Cell26, 323–334 (2007). [PubMed] [Google Scholar]
  38. Rossiello F et al. DNA damage response inhibition at dysfunctional telomeres by modulation of telomeric DNA damage response RNAs. Nat. Commun8, 13980 (2017). [PMC free article] [PubMed] [Google Scholar]
  39. Halazonetis TD, Gorgoulis VG & Bartek J An oncogene-induced DNA damage model for cancer development. Science319, 1352–1355 (2008). [PubMed] [Google Scholar]
  40. Alimonti A et al. A novel type of cellular senescence that can be enhanced in mouse models and human tumor xenografts to suppress prostate tumorigenesis. J. Clin. Invest120, 681–693 (2010). [PMC free article] [PubMed] [Google Scholar]
  41. Parisotto M et al. PTEN deletion in luminal cells of mature prostate induces replication stress and senescence in vivo. J. Exp. Med215, 1749–1763 (2018). [PMC free article] [PubMed] [Google Scholar]
  42. Astle MV et al. AKT induces senescence in human cells via mTORC1 and p53 in the absence of DNA damage: implications for targeting mTOR during malignancy. Oncogene31, 1949–1962 (2012). [PMC free article] [PubMed] [Google Scholar]
  43. Chan KT et al. A functional genetic screen defines the AKT-induced senescence signaling network. Cell Death Differ. 27, 725–741 (2020). [PMC free article] [PubMed] [Google Scholar]
  44. Suram A et al. Oncogene-induced telomere dysfunction enforces cellular senescence in human cancer precursor lesions. EMBO J. 31, 2839–2851 (2012). [PMC free article] [PubMed] [Google Scholar]
  45. Ogrunc M et al. Oncogene-induced reactive oxygen species fuel hyperproliferation and DNA damage response activation. Cell Death Differ. 21, 998–1012 (2014). [PMC free article] [PubMed] [Google Scholar]
  46. Chapman J, Fielder E & Passos JF Mitochondrial dysfunction and cell senescence: deciphering a complex relationship. FEBS Lett. 593, 1566–1579 (2019). [PubMed] [Google Scholar]
  47. Wiley CD et al. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 23, 303–314 (2016). [PMC free article] [PubMed] [Google Scholar]
  48. Correia-Melo C et al. Mitochondria are required for pro-ageing features of the senescent phenotype. EMBO J. 35, 724–742 (2016). [PMC free article] [PubMed] [Google Scholar]
  49. Narita M et al. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell113, 703–716 (2003). [PubMed] [Google Scholar]
  50. Zhang R et al. Formation of MacroH2A-containing senescence-associated heterochromatin foci and senescence driven by ASF1a and HIRA. Dev. Cell8, 19–30 (2005). [PubMed] [Google Scholar]
  51. Chandra T et al. Independence of repressive histone marks and chromatin compaction during senescent heterochromatic layer formation. Mol. Cell47, 203–214 (2012). [PMC free article] [PubMed] [Google Scholar]
  52. Di Micco R et al. Interplay between oncogene-induced DNA damage response and heterochromatin in senescence and cancer. Nat. Cell Biol13, 292–302 (2011). [PMC free article] [PubMed] [Google Scholar]
  53. Zhang R, Chen W & Adams PD Molecular dissection of formation of senescence-associated heterochromatin foci. Mol. Cell. Biol27, 2343–2358 (2007). [PMC free article] [PubMed] [Google Scholar]
  54. Sadaie M et al. Redistribution of the lamin B1 genomic binding profile affects rearrangement of heterochromatic domains and SAHF formation during senescence. Genes Dev. 27, 1800–1808 (2013). [PMC free article] [PubMed] [Google Scholar]
  55. Samaraweera L, Adomako A, Rodriguez-Gabin A & McDaid HM A novel indication for panobinostat as a senolytic drug in NSCLC and HNSCC. Sci. Rep7, 1900 (2017). [PMC free article] [PubMed] [Google Scholar]
  56. Munro J, Barr NI, Ireland H, Morrison V & Parkinson EK Histone deacetylase inhibitors induce a senescence-like state in human cells by a p16-dependent mechanism that is independent of a mitotic clock. Exp. Cell. Res295, 525–538 (2004). [PubMed] [Google Scholar]
  57. Swanson EC, Manning B, Zhang H & Lawrence JB Higher-order unfolding of satellite heterochromatin is a consistent and early event in cell senescence. J. Cell Biol203, 929–942 (2013). [PMC free article] [PubMed] [Google Scholar]
  58. Ivanov A et al. Lysosome-mediated processing of chromatin in senescence. J. Cell Biol202, 129–143 (2013). [PMC free article] [PubMed] [Google Scholar]
  59. Dou Z et al. Autophagy mediates degradation of nuclear lamina. Nature527, 105–109 (2015). [PMC free article] [PubMed] [Google Scholar]
  60. Dou Z et al. Cytoplasmic chromatin triggers inflammation in senescence and cancer. Nature550, 402–406 (2017). [PMC free article] [PubMed] [Google Scholar]This study demonstrates that cytoplasmic chromatin in senescent and cancer cells activates the innate immunity through the cGAS–STING pathway.
  61. Tsantoulis PK et al. Oncogene-induced replication stress preferentially targets common fragile sites in preneoplastic lesions. A genome-wide study. Oncogene27, 3256–3264 (2008). [PubMed] [Google Scholar]
  62. De Cecco M et al. L1 drives IFN in senescent cells and promotes age-associated inflammation. Nature566, 73–78 (2019). [PMC free article] [PubMed] [Google Scholar]
  63. Vizioli MG et al. Mitochondria-to-nucleus retrograde signaling drives formation of cytoplasmic chromatin and inflammation in senescence. Genes Dev. 34, 428–445 (2020). [PMC free article] [PubMed] [Google Scholar]
  64. Cruickshanks HA et al. Senescent cells harbour features of the cancer epigenome. Nat. Cell Biol15, 1495–1506 (2013). [PMC free article] [PubMed] [Google Scholar]
  65. Xie W et al. DNA methylation patterns separate senescence from transformation potential and indicate cancer risk. Cancer Cell33, 309–321 e305 (2018). [PMC free article] [PubMed] [Google Scholar]
  66. Parry AJ et al. NOTCH-mediated non-cell autonomous regulation of chromatin structure during senescence. Nat. Commun9, 1840 (2018). [PMC free article] [PubMed] [Google Scholar]
  67. De Cecco M et al. Genomes of replicatively senescent cells undergo global epigenetic changes leading to gene silencing and activation of transposable elements. Aging Cell12, 247–256 (2013). [PMC free article] [PubMed] [Google Scholar]
  68. Shah PP et al. Lamin B1 depletion in senescent cells triggers large-scale changes in gene expression and the chromatin landscape. Genes Dev. 27, 1787–1799 (2013). [PMC free article] [PubMed] [Google Scholar]
  69. Freund A, Laberge RM, Demaria M & Campisi J Lamin B1 loss is a senescence-associated biomarker. Mol. Biol. Cell23, 2066–2075 (2012). [PMC free article] [PubMed] [Google Scholar]
  70. Shimi T et al. The role of nuclear lamin B1 in cell proliferation and senescence. Genes Dev. 25, 2579–2593 (2011). [PMC free article] [PubMed] [Google Scholar]
  71. Tasdemir N et al. BRD4 connects enhancer remodeling to senescence immune surveillance. Cancer Discov. 6, 612–629 (2016). [PMC free article] [PubMed] [Google Scholar]
  72. Sen P et al. Histone acetyltransferase p300 induces de novo super-enhancers to drive cellular senescence. Mol. Cell73, 684–698 e688 (2019). [PMC free article] [PubMed] [Google Scholar]
  73. Acosta JC et al. A complex secretory program orchestrated by the inflammasome controls paracrine senescence. Nat. Cell Biol15, 978–990 (2013). [PMC free article] [PubMed] [Google Scholar]
  74. Kuilman T et al. Oncogene-induced senescence relayed by an interleukin-dependent inflammatory network. Cell133, 1019–1031 (2008). [PubMed] [Google Scholar]
  75. Coppe JP et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. 6, 2853–2868 (2008). [PMC free article] [PubMed] [Google Scholar]
  76. Coppe JP, Desprez PY, Krtolica A & Campisi J The senescence-associated secretory phenotype: the dark side of tumor suppression. Annu. Rev. Pathol5, 99–118 (2010). [PMC free article] [PubMed] [Google Scholar]
  77. Coppe JP et al. A human-like senescence-associated secretory phenotype is conserved in mouse cells dependent on physiological oxygen. PLoS ONE5, e9188 (2010). [PMC free article] [PubMed] [Google Scholar]
  78. Eren M et al. PAI-1-regulated extracellular proteolysis governs senescence and survival in Klotho mice. Proc. Natl Acad. Sci. USA111, 7090–7095 (2014). [PMC free article] [PubMed] [Google Scholar]
  79. Ozcan S et al. Unbiased analysis of senescence associated secretory phenotype (SASP) to identify common components following different genotoxic stresses. Aging8, 1316–1329 (2016). [PMC free article] [PubMed] [Google Scholar]
  80. Basisty N et al. A proteomic atlas of senescence-associated secretomes for aging biomarker development. PLoS Biol. 18, e3000599 (2020). [PMC free article] [PubMed] [Google Scholar]
  81. Faget DV, Ren Q & Stewart SA Unmasking senescence: context-dependent effects of SASP in cancer. Nat. Rev. Cancer19, 439–453 (2019). [PubMed] [Google Scholar]
  82. Tanaka T et al. Plasma proteomic signature of age in healthy humans. Aging Cell17, e12799 (2018). [PMC free article] [PubMed] [Google Scholar]
  83. Borghesan M et al. Small extracellular vesicles are key regulators of non-cell autonomous intercellular communication in senescence via the interferon protein IFITM3. Cell Rep. 27, 3956–3971 e3956 (2019). [PMC free article] [PubMed] [Google Scholar]
  84. Jakhar R & Crasta K Exosomes as emerging pro-tumorigenic mediators of the senescence-associated secretory phenotype. Int. J. Mol. Sci10.3390/ijms20102547 (2019). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  85. Coppe JP et al. Tumor suppressor and aging biomarker p16(INK4a) induces cellular senescence without the associated inflammatory secretory phenotype. J. Biol. Chem286, 36396–36403 (2011). [PMC free article] [PubMed] [Google Scholar]
  86. Rodier F et al. Persistent DNA damage signalling triggers senescence-associated inflammatory cytokine secretion. Nat. Cell Biol11, 973–979 (2009). [PMC free article] [PubMed] [Google Scholar]This study demonstrates that persistent DDR activation controls SASP induction in senescent cells.
  87. Chen H et al. MacroH2A1 and ATM play opposing roles in paracrine senescence and the senescence-associated secretory phenotype. Mol. Cell59, 719–731 (2015). [PMC free article] [PubMed] [Google Scholar]
  88. Freund A, Patil CK & Campisi J p38MAPK is a novel DNA damage response-independent regulator of the senescence-associated secretory phenotype. EMBO J. 30, 1536–1548 (2011). [PMC free article] [PubMed] [Google Scholar]
  89. White RR, Vijg J & Do DNA Double-strand breaks drive aging? Mol. Cell63, 729–738 (2016). [PMC free article] [PubMed] [Google Scholar]
  90. Ou HL & Schumacher B DNA damage responses and p53 in the aging process. Blood131, 488–495 (2018). [PMC free article] [PubMed] [Google Scholar]
  91. Gorbunova V & Seluanov A DNA double strand break repair, aging and the chromatin connection. Mutat. Res788, 2–6 (2016). [PMC free article] [PubMed] [Google Scholar]
  92. Shmulevich R & Krizhanovsky V Cell senescence, DNA damage, and metabolism. Antioxid. Redox Signal10.1089/ars.2020.8043 (2020). [PubMed] [CrossRef] [Google Scholar]
  93. Acosta JC et al. Chemokine signaling via the CXCR2 receptor reinforces senescence. Cell133, 1006–1018 (2008). [PubMed] [Google Scholar]
  94. Hoare M et al. NOTCH1 mediates a switch between two distinct secretomes during senescence. Nat. Cell Biol18, 979–992 (2016). [PMC free article] [PubMed] [Google Scholar]
  95. Kang C et al. The DNA damage response induces inflammation and senescence by inhibiting autophagy of GATA4. Science349, aaa5612 (2015). [PMC free article] [PubMed] [Google Scholar]
  96. Mazzucco AE et al. Genetic interrogation of replicative senescence uncovers a dual role for USP28 in coordinating the p53 and GATA4 branches of the senescence program. Genes Dev. 31, 1933–1938 (2017). [PMC free article] [PubMed] [Google Scholar]
  97. Orjalo AV, Bhaumik D, Gengler BK, Scott GK & Campisi J Cell surface-bound IL-1alpha is an upstream regulator of the senescence-associated IL-6/IL-8 cytokine network. Proc. Natl Acad. Sci. USA106, 17031–17036 (2009). [PMC free article] [PubMed] [Google Scholar]
  98. Toso A et al. Enhancing chemotherapy efficacy in Pten-deficient prostate tumors by activating the senescence-associated antitumor immunity. Cell Rep. 9, 75–89 (2014). [PubMed] [Google Scholar]
  99. Xu M et al. JAK inhibition alleviates the cellular senescence-associated secretory phenotype and frailty in old age. Proc. Natl Acad. Sci. USA112, E6301–E6310 (2015). [PMC free article] [PubMed] [Google Scholar]
  100. Hernandez-Segura A et al. Unmasking transcriptional heterogeneity in senescent cells. Curr. Biol27, 2652–2660 e2654 (2017). [PMC free article] [PubMed] [Google Scholar]
  101. Vermezovic J et al. Notch is a direct negative regulator of the DNA-damage response. Nat. Struct. Mol. Biol22, 417–424 (2015). [PubMed] [Google Scholar]
  102. Adamowicz M, Vermezovic J & d’Adda di Fagagna F NOTCH1 inhibits activation of ATM by impairing the formation of an ATM-FOXO3a-KAT5/Tip60 complex. Cell Rep. 16, 2068–2076 (2016). [PMC free article] [PubMed] [Google Scholar]
  103. Stathis A & Bertoni F BET proteins as targets for anticancer treatment. Cancer Discov. 8, 24–36 (2018). [PubMed] [Google Scholar]
  104. Wakita M et al. A BET family protein degrader provokes senolysis by targeting NHEJ and autophagy in senescent cells. Nat. Commun11, 1935 (2020). [PMC free article] [PubMed] [Google Scholar]
  105. Wang S et al. BRD4 inhibitors block telomere elongation. Nucleic Acids Res. 45, 8403–8410 (2017). [PMC free article] [PubMed] [Google Scholar]
  106. Ito T, Teo YV, Evans SA, Neretti N & Sedivy JM Regulation of cellular senescence by polycomb chromatin modifiers through distinct DNA damage-and histone methylation-dependent pathways. Cell Rep. 22, 3480–3492 (2018). [PMC free article] [PubMed] [Google Scholar]
  107. Capell BC et al. MLL1 is essential for the senescence-associated secretory phenotype. Genes Dev. 30, 321–336 (2016). [PMC free article] [PubMed] [Google Scholar]
  108. Aird KM et al. HMGB2 orchestrates the chromatin landscape of senescence-associated secretory phenotype gene loci. J. Cell Biol215, 325–334 (2016). [PMC free article] [PubMed] [Google Scholar]
  109. Huang J et al. DAMPs, ageing, and cancer: the ‘DAMP hypothesis’. Ageing Res. Rev24, 3–16 (2015). [PMC free article] [PubMed] [Google Scholar]
  110. Davalos AR et al. p53-dependent release of alarmin HMGB1 is a central mediator of senescent phenotypes. J. Cell Biol201, 613–629 (2013). [PMC free article] [PubMed] [Google Scholar]
  111. Boumendil C, Hari P, Olsen KCF, Acosta JC & Bickmore WA Nuclear pore density controls heterochromatin reorganization during senescence. Genes Dev. 33, 144–149 (2019). [PMC free article] [PubMed] [Google Scholar]
  112. Gluck S et al. Innate immune sensing of cytosolic chromatin fragments through cGAS promotes senescence. Nat. Cell Biol19, 1061–1070 (2017). [PMC free article] [PubMed] [Google Scholar]
  113. Yang H, Wang H, Ren J, Chen Q & Chen ZJ cGAS is essential for cellular senescence. Proc. Natl Acad. Sci. USA114, E4612–E4620 (2017). [PMC free article] [PubMed] [Google Scholar]
  114. Hopfner KP & Hornung V Molecular mechanisms and cellular functions of cGAS-STING signalling. Nat. Rev. Mol. Cell Biol21, 501–521 (2020). [PubMed] [Google Scholar]
  115. Takahashi A et al. Downregulation of cytoplasmic DNases is implicated in cytoplasmic DNA accumulation and SASP in senescent cells. Nat. Commun9, 1249 (2018). [PMC free article] [PubMed] [Google Scholar]
  116. Haag SM et al. Targeting STING with covalent small-molecule inhibitors. Nature559, 269–273 (2018). [PubMed] [Google Scholar]
  117. Broz P & Dixit VM Inflammasomes: mechanism of assembly, regulation and signalling. Nat. Rev. Immunol16, 407–420 (2016). [PubMed] [Google Scholar]
  118. Hari P et al. The innate immune sensor Toll-like receptor 2 controls the senescence-associated secretory phenotype. Sci. Adv5, eaaw0254 (2019). [PMC free article] [PubMed] [Google Scholar]
  119. Simon M et al. LINE1 derepression in aged wild-type and SIRT6-deficient mice drives inflammation. Cell Metab. 29, 871–885 e875 (2019). [PMC free article] [PubMed] [Google Scholar]
  120. Warren LA & Rossi DJ Stem cells and aging in the hematopoietic system. Mech. Ageing Dev130, 46–53 (2009). [PMC free article] [PubMed] [Google Scholar]
  121. Rossi DJ et al. Deficiencies in DNA damage repair limit the function of haematopoietic stem cells with age. Nature447, 725–729 (2007). [PubMed] [Google Scholar]
  122. McNeely T, Leone M, Yanai H & Beerman I DNA damage in aging, the stem cell perspective. Hum. Genet10.1007/s00439-019-02047-z (2019). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  123. Sperka T, Wang J & Rudolph KL DNA damage checkpoints in stem cells, ageing and cancer. Nat. Rev. Mol. Cell Biol13, 579–590 (2012). [PubMed] [Google Scholar]
  124. Inomata K et al. Genotoxic stress abrogates renewal of melanocyte stem cells by triggering their differentiation. Cell137, 1088–1099 (2009). [PubMed] [Google Scholar]This study demonstrates that DNA damage can induce stem cell differentiation.
  125. Wang J et al. A differentiation checkpoint limits hematopoietic stem cell self-renewal in response to DNA damage. Cell158, 1444 (2014). [PubMed] [Google Scholar]
  126. Schiroli G et al. Precise gene editing preserves hematopoietic stem cell function following transient p53-mediated DNA damage response. Cell Stem Cell24, 551–565 e558 (2019). [PMC free article] [PubMed] [Google Scholar]
  127. Flach J et al. Replication stress is a potent driver of functional decline in ageing haematopoietic stem cells. Nature512, 198–202 (2014). [PMC free article] [PubMed] [Google Scholar]
  128. Schneider L et al. DNA damage in mammalian neural stem cells leads to astrocytic differentiation mediated by BMP2 signaling through JAK-STAT. Stem Cell Rep. 1, 123–138 (2013). [PMC free article] [PubMed] [Google Scholar]
  129. Zou Y et al. Responses of human embryonic stem cells and their differentiated progeny to ionizing radiation. Biochem. Biophys. Res. Commun426, 100–105 (2012). [PMC free article] [PubMed] [Google Scholar]
  130. Li M et al. Distinct regulatory mechanisms and functions for p53-activated and p53-repressed DNA damage response genes in embryonic stem cells. Mol. Cell46, 30–42 (2012). [PMC free article] [PubMed] [Google Scholar]
  131. Storer M et al. Senescence is a developmental mechanism that contributes to embryonic growth and patterning. Cell155, 1119–1130 (2013). [PubMed] [Google Scholar]
  132. Munoz-Espin D et al. Programmed cell senescence during mammalian embryonic development. Cell155, 1104–1118 (2013). [PubMed] [Google Scholar]
  133. Sapieha P & Mallette FA Cellular senescence in postmitotic cells: beyond growth arrest. Trends Cell Biol. 28, 595–607 (2018). [PubMed] [Google Scholar]
  134. Jurk D et al. Postmitotic neurons develop a p21-dependent senescence-like phenotype driven by a DNA damage response. Aging Cell11, 996–1004 (2012). [PMC free article] [PubMed] [Google Scholar]
  135. Ogrodnik M et al. Obesity-induced cellular senescence drives anxiety and impairs neurogenesis. Cell Metab. 29, 1061–1077.e1068 (2019). [PMC free article] [PubMed] [Google Scholar]
  136. Musi N et al. Tau protein aggregation is associated with cellular senescence in the brain. Aging Cell17, e12840 (2018). [PMC free article] [PubMed] [Google Scholar]
  137. Oubaha M et al. Senescence-associated secretory phenotype contributes to pathological angiogenesis in retinopathy. Sci. Transl Med8, 362ra144 (2016). [PubMed] [Google Scholar]
  138. Minamino T et al. A crucial role for adipose tissue p53 in the regulation of insulin resistance. Nat. Med15, 1082–1087 (2009). [PubMed] [Google Scholar]
  139. Farr JN et al. Targeting cellular senescence prevents age-related bone loss in mice. Nat. Med23, 1072–1079 (2017). [PMC free article] [PubMed] [Google Scholar]
  140. Anderson R et al. Length-independent telomere damage drives post-mitotic cardiomyocyte senescence. EMBO J. 10.15252/embj.2018100492 (2019). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  141. Burton DG & Krizhanovsky V Physiological and pathological consequences of cellular senescence. Cell. Mol. Life Sci71, 4373–4386 (2014). [PMC free article] [PubMed] [Google Scholar]
  142. Chuprin A et al. Cell fusion induced by ERVWE1 or measles virus causes cellular senescence. Genes Dev. 27, 2356–2366 (2013). [PMC free article] [PubMed] [Google Scholar]
  143. Gal H et al. Molecular pathways of senescence regulate placental structure and function. EMBO J. 38, e100849 (2019). [PMC free article] [PubMed] [Google Scholar]
  144. Davaapil H, Brockes JP & Yun MH Conserved and novel functions of programmed cellular senescence during vertebrate development. Development144, 106–114 (2017). [PMC free article] [PubMed] [Google Scholar]
  145. Krizhanovsky V et al. Senescence of activated stellate cells limits liver fibrosis. Cell134, 657–667 (2008). [PMC free article] [PubMed] [Google Scholar]This study demonstrates that immune cell targeting of senescent cells limits dysfunction in the liver.
  146. Kim KH, Chen CC, Monzon RI & Lau LF Matricellular protein CCN1 promotes regression of liver fibrosis through induction of cellular senescence in hepatic myofibroblasts. Mol. Cell. Biol33, 2078–2090 (2013). [PMC free article] [PubMed] [Google Scholar]
  147. Jun JI & Lau LF The matricellular protein CCN1 induces fibroblast senescence and restricts fibrosis in cutaneous wound healing. Nat. Cell Biol12, 676–685 (2010). [PMC free article] [PubMed] [Google Scholar]
  148. Fitzner B et al. Senescence determines the fate of activated rat pancreatic stellate cells. J. Cell Mol. Med16, 2620–2630 (2012). [PMC free article] [PubMed] [Google Scholar]
  149. Da Silva-Álvarez S et al. Cell senescence contributes to tissue regeneration in zebrafish. Aging Cell19, e13052 (2020). [PMC free article] [PubMed] [Google Scholar]
  150. Yun MH, Davaapil H & Brockes JP Recurrent turnover of senescent cells during regeneration of a complex structure. eLife10.7554/eLife.05505 (2015). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  151. Sagiv A & Krizhanovsky V Immunosurveillance of senescent cells: the bright side of the senescence program. Biogerontology14, 617–628 (2013). [PubMed] [Google Scholar]
  152. Freund A, Orjalo AV, Desprez PY & Campisi J Inflammatory networks during cellular senescence: causes and consequences. Trends Mol. Med16, 238–246 (2010). [PMC free article] [PubMed] [Google Scholar]
  153. Lujambio A et al. Non-cell-autonomous tumor suppression by p53. Cell153, 449–460 (2013). [PMC free article] [PubMed] [Google Scholar]
  154. Soriani A et al. ATM-ATR-dependent up-regulation of DNAM-1 and NKG2D ligands on multiple myeloma cells by therapeutic agents results in enhanced NK-cell susceptibility and is associated with a senescent phenotype. Blood113, 3503–3511 (2009). [PubMed] [Google Scholar]
  155. Xue W et al. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature445, 656–660 (2007). [PMC free article] [PubMed] [Google Scholar]
  156. Kang TW et al. Senescence surveillance of pre-malignant hepatocytes limits liver cancer development. Nature479, 547–551 (2011). [PubMed] [Google Scholar]
  157. Sagiv A et al. NKG2D ligands mediate immunosurveillance of senescent cells. Aging8, 328–344 (2016). [PMC free article] [PubMed] [Google Scholar]
  158. Biran A et al. Senescent cells communicate via intercellular protein transfer. Genes Dev. 29, 791–802 (2015). [PMC free article] [PubMed] [Google Scholar]
  159. Ruscetti M et al. NK cell-mediated cytotoxicity contributes to tumor control by a cytostatic drug combination. Science362, 1416–1422 (2018). [PMC free article] [PubMed] [Google Scholar]
  160. Di Mitri D et al. Tumour-infiltrating Gr-1+ myeloid cells antagonize senescence in cancer. Nature515, 134–137 (2014). [PubMed] [Google Scholar]
  161. Eggert T et al. Distinct functions of senescence-associated immune responses in liver tumor surveillance and tumor progression. Cancer Cell30, 533–547 (2016). [PMC free article] [PubMed] [Google Scholar]
  162. Yoshimoto S et al. Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome. Nature499, 97–101 (2013). [PubMed] [Google Scholar]
  163. Gonzalez-Meljem JM et al. Stem cell senescence drives age-attenuated induction of pituitary tumours in mouse models of paediatric craniopharyngioma. Nat. Commun8, 1819 (2017). [PMC free article] [PubMed] [Google Scholar]
  164. Lau L, Porciuncula A, Yu A, Iwakura Y & David G Uncoupling the senescence-associated secretory phenotype from cell cycle exit via interleukin-1 inactivation unveils its protumorigenic role. Mol. Cell. Biol39, e00586–18 (2019). [PMC free article] [PubMed] [Google Scholar]
  165. Demaria M et al. Cellular senescence promotes adverse effects of chemotherapy and cancer relapse. Cancer Discov. 7, 165–176 (2017). [PMC free article] [PubMed] [Google Scholar]
  166. Franceschi C et al. Inflamm-aging. An evolutionary perspective on immunosenescence. Ann. N. Y. Acad. Sci908, 244–254 (2000). [PubMed] [Google Scholar]
  167. Jeck WR, Siebold AP & Sharpless NE Review: a meta-analysis of GWAS and age-associated diseases. Aging Cell11, 727–731 (2012). [PMC free article] [PubMed] [Google Scholar]
  168. Melzer D Genetic polymorphisms and human aging: association studies deliver. Rejuvenation Res. 11, 523–526 (2008). [PubMed] [Google Scholar]
  169. Melzer D et al. A common variant of the p16(INK4a) genetic region is associated with physical function in older people. Mech. Ageing Dev128, 370–377 (2007). [PMC free article] [PubMed] [Google Scholar]
  170. Johnson SC, Dong X, Vijg J & Suh Y Genetic evidence for common pathways in human age-related diseases. Aging Cell14, 809–817 (2015). [PMC free article] [PubMed] [Google Scholar]
  171. D’Mello MJ et al. Association between shortened leukocyte telomere length and cardiometabolic outcomes: systematic review and meta-analysis. Circ. Cardiovasc. Genet8, 82–90 (2015). [PubMed] [Google Scholar]
  172. Benetos A et al. Tracking and fixed ranking of leukocyte telomere length across the adult life course. Aging Cell12, 615–621 (2013). [PMC free article] [PubMed] [Google Scholar]
  173. Fabbri E et al. Aging and the burden of multimorbidity: associations with inflammatory and anabolic hormonal biomarkers. J. Gerontol. A Biol. Sci. Med. Sci70, 63–70 (2015). [PMC free article] [PubMed] [Google Scholar]
  174. Xu M et al. Senolytics improve physical function and increase lifespan in old age. Nat. Med24, 1246–1256 (2018). [PMC free article] [PubMed] [Google Scholar]This study demonstrates that transplanting senescent cells in young mice causes persistent physical dysfunction that can be alleviated by senolytic drugs.
  175. Childs BG et al. Senescent intimal foam cells are deleterious at all stages of atherosclerosis. Science354, 472–477 (2016). [PMC free article] [PubMed] [Google Scholar]
  176. Holdt LM et al. Expression of Chr9p21 genes CDKN2B (p15(INK4b)), CDKN2A (p16(INK4a), p14(ARF)) and MTAP in human atherosclerotic plaque. Atherosclerosis214, 264–270 (2011). [PubMed] [Google Scholar]
  177. Cavalli G, Biavasco R, Borgiani B & Dagna L Oncogene-induced senescence as a new mechanism of disease: the paradigm of erdheim-chester disease. Front. Immunol5, 281 (2014). [PMC free article] [PubMed] [Google Scholar]
  178. Cangi MG et al. BRAFV600E-mutation is invariably present and associated to oncogene-induced senescence in Erdheim-Chester disease. Ann. Rheum. Dis74, 1596–1602 (2015). [PubMed] [Google Scholar]
  179. Sousa-Victor P, Garcia-Prat L, Serrano AL, Perdiguero E & Munoz-Canoves P Muscle stem cell aging: regulation and rejuvenation. Trends Endocrinol. Metab26, 287–296 (2015). [PubMed] [Google Scholar]
  180. Gnani D et al. An early-senescence state in aged mesenchymal stromal cells contributes to hematopoietic stem and progenitor cell clonogenic impairment through the activation of a pro-inflammatory program. Aging Cell18, e12933 (2019). [PMC free article] [PubMed] [Google Scholar]
  181. Krizhanovsky V & Lowe SW Stem cells: The promises and perils of p53. Nature460, 1085–1086 (2009). [PMC free article] [PubMed] [Google Scholar]
  182. Banito A et al. Senescence impairs successful reprogramming to pluripotent stem cells. Genes Dev. 23, 2134–2139 (2009). [PMC free article] [PubMed] [Google Scholar]
  183. Li H et al. The Ink4/Arf locus is a barrier for iPS cell reprogramming. Nature460, 1136–1139 (2009). [PMC free article] [PubMed] [Google Scholar]
  184. Mosteiro L et al. Tissue damage and senescence provide critical signals for cellular reprogramming in vivo. Science10.1126/science.aaf4445 (2016). [PubMed] [CrossRef] [Google Scholar]
  185. Wajapeyee N, Serra RW, Zhu X, Mahalingam M & Green MR Oncogenic BRAF induces senescence and apoptosis through pathways mediated by the secreted protein IGFBP7. Cell132, 363–374 (2008). [PMC free article] [PubMed] [Google Scholar]
  186. Takasugi M et al. Small extracellular vesicles secreted from senescent cells promote cancer cell proliferation through EphA2. Nat. Commun8, 15729 (2017). [PMC free article] [PubMed] [Google Scholar]
  187. Katlinskaya YV et al. Suppression of type I interferon signaling overcomes oncogene-induced senescence and mediates melanoma development and progression. Cell Rep. 15, 171–180 (2016). [PMC free article] [PubMed] [Google Scholar]
  188. Bird TG et al. TGFbeta inhibition restores a regenerative response in acute liver injury by suppressing paracrine senescence. Sci. Transl Med10.1126/scitranslmed.aan1230 (2018). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  189. Ritschka B et al. The senescence-associated secretory phenotype induces cellular plasticity and tissue regeneration. Genes Dev. 31, 172–183 (2017). [PMC free article] [PubMed] [Google Scholar]
  190. Chiche A et al. Injury-induced senescence enables in vivo reprogramming in skeletal muscle. Cell Stem Cell20, 407–414.e404 (2017). [PubMed] [Google Scholar]
  191. Baker DJ et al. BubR1 insufficiency causes early onset of aging-associated phenotypes and infertility in mice. Nat. Genet36, 744–749 (2004). [PubMed] [Google Scholar]
  192. Baker DJ, Jin F & van Deursen JM The yin and yang of the Cdkn2a locus in senescence and aging. Cell Cycle7, 2795–2802 (2008). [PMC free article] [PubMed] [Google Scholar]
  193. Baker DJ et al. Opposing roles for p16Ink4a and p19Arf in senescence and ageing caused by BubR1 insufficiency. Nat. Cell Biol10, 825–836 (2008). [PMC free article] [PubMed] [Google Scholar]
  194. Baker DJ et al. Naturally occurring p16(Ink4a)-positive cells shorten healthy lifespan. Nature530, 184–189 (2016). [PMC free article] [PubMed] [Google Scholar]
  195. Laberge RM et al. Mitochondrial DNA damage induces apoptosis in senescent cells. Cell Death Dis. 4, e727 (2013). [PMC free article] [PubMed] [Google Scholar]
  196. Chinta SJ et al. Cellular senescence is induced by the environmental neurotoxin paraquat and contributes to neuropathology linked to Parkinson’s disease. Cell Rep. 22, 930–940 (2018). [PMC free article] [PubMed] [Google Scholar]
  197. Bussian TJ et al. Clearance of senescent glial cells prevents tau-dependent pathology and cognitive decline. Nature562, 578–582 (2018). [PMC free article] [PubMed] [Google Scholar]
  198. Zhang P et al. Senolytic therapy alleviates Aβ-associated oligodendrocyte progenitor cell senescence and cognitive deficits in an Alzheimer’s disease model. Nat. Neurosci22, 719–728 (2019). [PMC free article] [PubMed] [Google Scholar]
  199. Schafer MJ et al. Cellular senescence mediates fibrotic pulmonary disease. Nat. Commun8, 14532 (2017). [PMC free article] [PubMed] [Google Scholar]
  200. Sagiv A et al. p53 in bronchial club cells facilitates chronic lung inflammation by promoting senescence. Cell Rep. 22, 3468–3479 (2018). [PubMed] [Google Scholar]
  201. Jeon OH et al. Local clearance of senescent cells attenuates the development of post-traumatic osteoarthritis and creates a pro-regenerative environment. Nat. Med23, 775–781 (2017). [PMC free article] [PubMed] [Google Scholar]
  202. Zhu Y et al. New agents that target senescent cells: the flavone, fisetin, and the BCL-X(L) inhibitors, A1331852 and A1155463. Aging9, 955–963 (2017). [PMC free article] [PubMed] [Google Scholar]
  203. Guerrero A et al. Cardiac glycosides are broad-spectrum senolytics. Nat. Metab1, 1074–1088 (2019). [PMC free article] [PubMed] [Google Scholar]
  204. Triana-Martinez F et al. Identification and characterization of cardiac glycosides as senolytic compounds. Nat. Commun10, 4731 (2019). [PMC free article] [PubMed] [Google Scholar]
  205. Li W, He Y, Zhang R, Zheng G & Zhou D The curcumin analog EF24 is a novel senolytic agent. Aging11, 771–782 (2019). [PMC free article] [PubMed] [Google Scholar]
  206. Baar MP et al. Targeted apoptosis of senescent cells restores tissue homeostasis in response to chemotoxicity and aging. Cell169, 132–147 e116 (2017). [PMC free article] [PubMed] [Google Scholar]
  207. Zhu Y et al. Inflammation and the depot-specific secretome of human preadipocytes. Obesity23, 989–999 (2015). [PMC free article] [PubMed] [Google Scholar]
  208. Roos CM et al. Chronic senolytic treatment alleviates established vasomotor dysfunction in aged or atherosclerotic mice. Aging Cell15, 973–977 (2016). [PMC free article] [PubMed] [Google Scholar]
  209. Yousefzadeh MJ et al. Fisetin is a senotherapeutic that extends health and lifespan. EBioMedicine36, 18–28 (2018). [PMC free article] [PubMed] [Google Scholar]
  210. Zhu Y et al. The Achilles’ heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell14, 644–658 (2015). [PMC free article] [PubMed] [Google Scholar]
  211. Hickson LJ et al. Senolytics decrease senescent cells in humans: preliminary report from a clinical trial of dasatinib plus quercetin in individuals with diabetic kidney disease. EBioMedicine47, 446–456 (2019). [PMC free article] [PubMed] [Google Scholar]
  212. Justice JN et al. Senolytics in idiopathic pulmonary fibrosis: results from a first-in-human, open-label, pilot study. EBioMedicine40, 554–563 (2019). [PMC free article] [PubMed] [Google Scholar]
  213. Fuhrmann-Stroissnigg H, Niedernhofer LJ & Robbins PD Hsp90 inhibitors as senolytic drugs to extend healthy aging. Cell Cycle17, 1048–1055 (2018). [PMC free article] [PubMed] [Google Scholar]
  214. Fuhrmann-Stroissnigg H et al. Identification of HSP90 inhibitors as a novel class of senolytics. Nat. Commun8, 422 (2017). [PMC free article] [PubMed] [Google Scholar]
  215. Wang Y et al. Discovery of piperlongumine as a potential novel lead for the development of senolytic agents. Aging8, 2915–2926 (2016). [PMC free article] [PubMed] [Google Scholar]
  216. Liu X et al. Senolytic activity of piperlongumine analogues: synthesis and biological evaluation. Bioorg. Med. Chem26, 3925–3938 (2018). [PMC free article] [PubMed] [Google Scholar]
  217. Ozsvari B, Nuttall JR, Sotgia F & Lisanti MP Azithromycin and roxithromycin define a new family of “senolytic” drugs that target senescent human fibroblasts. Aging10, 3294–3307 (2018). [PMC free article] [PubMed] [Google Scholar]
  218. Munoz-Espin D et al. A versatile drug delivery system targeting senescent cells. EMBO Mol. Med10.15252/emmm.201809355 (2018). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  219. Guerrero A et al. Galactose-modified duocarmycin prodrugs as senolytics. Aging Cell19, e13133 (2020). [PMC free article] [PubMed] [Google Scholar]
  220. González-Gualda E et al. Galacto-conjugation of navitoclax as an efficient strategy to increase senolytic specificity and reduce platelet toxicity. Aging Cell19, e13142 (2020). [PMC free article] [PubMed] [Google Scholar]
  221. Ovadya Y & Krizhanovsky V Strategies targeting cellular senescence. J. Clin. Invest128, 1247–1254 (2018). [PMC free article] [PubMed] [Google Scholar]
  222. McMichael EL et al. IL-21 enhances natural killer cell response to cetuximab-coated pancreatic tumor cells. Clin. Cancer Res23, 489–502 (2017). [PMC free article] [PubMed] [Google Scholar]
  223. Brady J et al. The interactions of multiple cytokines control NK cell maturation. J. Immunol185, 6679–6688 (2010). [PubMed] [Google Scholar]
  224. Elpek KG, Rubinstein MP, Bellemare-Pelletier A, Goldrath AW & Turley SJ Mature natural killer cells with phenotypic and functional alterations accumulate upon sustained stimulation with IL-15/IL-15Ralpha complexes. Proc. Natl Acad. Sci. USA107, 21647–21652 (2010). [PMC free article] [PubMed] [Google Scholar]
  225. Rautela J & Huntington ND IL-15 signaling in NK cell cancer immunotherapy. Curr. Opin. Immunol44, 1–6 (2017). [PubMed] [Google Scholar]
  226. Amor C et al. Senolytic CAR T cells reverse senescence-associated pathologies. Nature583, 127–132 (2020). [PMC free article] [PubMed] [Google Scholar]This study demonstrates the feasibility of targeting senescent cells in a variety of conditions by chimeric antigen receptor T cell-mediated therapy.
  227. Frescas D et al. Senescent cells expose and secrete an oxidized form of membrane-bound vimentin as revealed by a natural polyreactive antibody. Proc. Natl Acad. Sci. USA114, E1668–E1677 (2017). [PMC free article] [PubMed] [Google Scholar]
  228. Kim KM et al. Identification of senescent cell surface targetable protein DPP4. Genes Dev. 31, 1529–1534 (2017). [PMC free article] [PubMed] [Google Scholar]
  229. Althubiti M et al. Characterization of novel markers of senescence and their prognostic potential in cancer. Cell Death Dis. 5, e1528 (2014). [PMC free article] [PubMed] [Google Scholar]
  230. Herranz N et al. mTOR regulates MAPKAPK2 translation to control the senescence-associated secretory phenotype. Nat. Cell Biol17, 1205–1217 (2015). [PMC free article] [PubMed] [Google Scholar]
  231. Laberge RM et al. MTOR regulates the pro-tumorigenic senescence-associated secretory phenotype by promoting IL1A translation. Nat. Cell Biol17, 1049–1061 (2015). [PMC free article] [PubMed] [Google Scholar]
  232. Zhang B et al. The senescence-associated secretory phenotype is potentiated by feedforward regulatory mechanisms involving Zscan4 and TAK1. Nat. Commun9, 1723 (2018). [PMC free article] [PubMed] [Google Scholar]
  233. Thapa RK et al. Progressive slowdown/prevention of cellular senescence by CD9-targeted delivery of rapamycin using lactose-wrapped calcium carbonate nanoparticles. Sci. Rep7, 43299 (2017). [PMC free article] [PubMed] [Google Scholar]
  234. Harrison DE et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature460, 392–395 (2009). [PMC free article] [PubMed] [Google Scholar]
  235. Moiseeva O et al. Metformin inhibits the senescence-associated secretory phenotype by interfering with IKK/NF-kappaB activation. Aging Cell12, 489–498 (2013). [PubMed] [Google Scholar]
  236. Maruthur NM et al. Diabetes medications as monotherapy or metformin-based combination therapy for type 2 diabetes: a systematic review and meta-analysis. Ann. Intern. Med164, 740–751 (2016). [PubMed] [Google Scholar]
  237. Noren Hooten N et al. Metformin-mediated increase in DICER1 regulates microRNA expression and cellular senescence. Aging Cell15, 572–581 (2016). [PMC free article] [PubMed] [Google Scholar]
  238. Lim H, Park H & Kim HP Effects of flavonoids on senescence-associated secretory phenotype formation from bleomycin-induced senescence in BJ fibroblasts. Biochem. Pharmacol96, 337–348 (2015). [PubMed] [Google Scholar]
  239. Nacarelli T et al. NAD+metabolism governs the proinflammatory senescence-associated secretome. Nat. Cell Biol 21, 397–407 (2019). [PMC free article] [PubMed] [Google Scholar]
  240. Liu S et al. Simvastatin suppresses breast cancer cell proliferation induced by senescent cells. Sci. Rep5, 17895 (2015). [PMC free article] [PubMed] [Google Scholar]
  241. Laberge RM et al. Glucocorticoids suppress selected components of the senescence-associated secretory phenotype. Aging Cell11, 569–578 (2012). [PMC free article] [PubMed] [Google Scholar]
  242. Cohen S et al. Treatment of rheumatoid arthritis with anakinra, a recombinant human interleukin-1 receptor antagonist, in combination with methotrexate: results of a twenty-four-week, multicenter, randomized, double-blind, placebo-controlled trial. Arthritis Rheum. 46, 614–624 (2002). [PubMed] [Google Scholar]
  243. Kuemmerle-Deschner JB et al. Canakinumab (ACZ885, a fully human IgG1 anti-IL-1β mAb) induces sustained remission in pediatric patients with cryopyrin-associated periodic syndrome (CAPS). Arthritis Res. Ther13, R34 (2011). [PMC free article] [PubMed] [Google Scholar]
  244. Hoffman HM et al. Efficacy and safety of rilonacept (interleukin-1 trap) in patients with cryopyrin-associated periodic syndromes: results from two sequential placebo-controlled studies. Arthritis Rheum. 58, 2443–2452 (2008). [PubMed] [Google Scholar]
  245. Klareskog L et al. Therapeutic effect of the combination of etanercept and methotrexate compared with each treatment alone in patients with rheumatoid arthritis: double-blind randomised controlled trial. Lancet363, 675–681 (2004). [PubMed] [Google Scholar]
  246. Jobanputra P, Barton P, Bryan S & Burls A The effectiveness of infliximab and etanercept for the treatment of rheumatoid arthritis: a systematic review and economic evaluation. Health Technol. Assess6, 1–110 (2002). [PubMed] [Google Scholar]
  247. Emery P et al. IL-6 receptor inhibition with tocilizumab improves treatment outcomes in patients with rheumatoid arthritis refractory to anti-tumour necrosis factor biologicals: results from a 24-week multicentre randomised placebo-controlled trial. Ann. Rheum. Dis67, 1516–1523 (2008). [PMC free article] [PubMed] [Google Scholar]
  248. van Rhee F et al. Siltuximab for multicentric Castleman’s disease: a randomised, double-blind, placebo-controlled trial. Lancet Oncol. 15, 966–974 (2014). [PubMed] [Google Scholar]
  249. Harrison DE et al. Acarbose, 17-α-estradiol, and nordihydroguaiaretic acid extend mouse lifespan preferentially in males. Aging Cell13, 273–282 (2014). [PMC free article] [PubMed] [Google Scholar]
  250. Michelini F et al. Damage-induced lncRNAs control the DNA damage response through interaction with DDRNAs at individual double-strand breaks. Nat. Cell Biol19, 1400–1411 (2017). [PMC free article] [PubMed] [Google Scholar]
  251. Aguado J et al. Inhibition of DNA damage response at telomeres improves the detrimental phenotypes of Hutchinson-Gilford progeria syndrome. Nat. Commun10, 4990 (2019). [PMC free article] [PubMed] [Google Scholar]
  252. Hall BM et al. Aging of mice is associated with p16(Ink4a)- and beta-galactosidase-positive macrophage accumulation that can be induced in young mice by senescent cells. Aging8, 1294–1315 (2016). [PMC free article] [PubMed] [Google Scholar]
  253. Liu JY et al. Cells exhibiting strong p16 (INK4a) promoter activation in vivo display features of senescence. Proc. Natl Acad. Sci. USA116, 2603–2611 (2019). [PMC free article] [PubMed] [Google Scholar]
  254. Grosse L et al. Defined p16(High) senescent cell types are indispensable for mouse healthspan. Cell Metab10.1016/j.cmet.2020.05.002 (2020). [PubMed] [CrossRef] [Google Scholar]
  255. Fontana L, Partridge L & Longo VD Extending healthy life span–from yeast to humans. Science328, 321–326 (2010). [PMC free article] [PubMed] [Google Scholar]
  256. Krishnamurthy J et al. Ink4a/Arf expression is a biomarker of aging. J. Clin. Invest114, 1299–1307 (2004). [PMC free article] [PubMed] [Google Scholar]
  257. Fontana L et al. The effects of graded caloric restriction: XII. Comparison of mouse to human impact on cellular senescence in the colon. Aging Cell17, e12746 (2018). [PMC free article] [PubMed] [Google Scholar]
  258. Wang C et al. Adult-onset, short-term dietary restriction reduces cell senescence in mice. Aging2, 555–566 (2010). [PMC free article] [PubMed] [Google Scholar]
  259. Satyanarayana A et al. Mitogen stimulation cooperates with telomere shortening to activate DNA damage responses and senescence signaling. Mol. Cell. Biol24, 5459–5474 (2004). [PMC free article] [PubMed] [Google Scholar]
  260. Martinez P & Blasco MA Telomere-driven diseases and telomere-targeting therapies. J. Cell Biol216, 875–887 (2017). [PMC free article] [PubMed] [Google Scholar]
  261. Salvador L et al. A natural product telomerase activator lengthens telomeres in humans: a randomized, double blind, and placebo controlled study. Rejuvenation Res. 19, 478–484 (2016). [PMC free article] [PubMed] [Google Scholar]
  262. Bar C, Huber N, Beier F & Blasco MA Therapeutic effect of androgen therapy in a mouse model of aplastic anemia produced by short telomeres. Haematologica100, 1267–1274 (2015). [PMC free article] [PubMed] [Google Scholar]
  263. Bernardes de Jesus B et al. Telomerase gene therapy in adult and old mice delays aging and increases longevity without increasing cancer. EMBO Mol. Med4, 691–704 (2012). [PMC free article] [PubMed] [Google Scholar]
  264. Alvarez D et al. IPF lung fibroblasts have a senescent phenotype. Am. J. Physiol. Lung Cell Mol. Physiol313, L1164–L1173 (2017). [PMC free article] [PubMed] [Google Scholar]
  265. Povedano JM et al. Therapeutic effects of telomerase in mice with pulmonary fibrosis induced by damage to the lungs and short telomeres. eLife10.7554/eLife.31299 (2018). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  266. Munoz-Lorente MA et al. AAV9-mediated telomerase activation does not accelerate tumorigenesis in the context of oncogenic K-Ras-induced lung cancer. PLoS Genet. 14, e1007562 (2018). [PMC free article] [PubMed] [Google Scholar]
  267. Conboy MJ, Conboy IM & Rando TA Heterochronic parabiosis: historical perspective and methodological considerations for studies of aging and longevity. Aging Cell12, 525–530 (2013). [PMC free article] [PubMed] [Google Scholar]
  268. Castellano JM et al. Human umbilical cord plasma proteins revitalize hippocampal function in aged mice. Nature544, 488–492 (2017). [PMC free article] [PubMed] [Google Scholar]
  269. Conboy IM & Rando TA Heterochronic parabiosis for the study of the effects of aging on stem cells and their niches. Cell Cycle11, 2260–2267 (2012). [PMC free article] [PubMed] [Google Scholar]
  270. Yousefzadeh MJ et al. Heterochronic parabiosis regulates the extent of cellular senescence in multiple tissues. Geroscience10.1007/s11357-020-00185-1 (2020). [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  271. US National Library of Medicine. ClinicalTrials.govhttps://clinicaltrials.gov/ct2/show/NCT04057872 (2019). [Google Scholar]
  272. US National Library of Medicine. ClinicalTrials.govhttps://www.clinicaltrials.gov/ct2/show/NCT03702920 (2018). [Google Scholar]
  273. Karin O, Agrawal A, Porat Z, Krizhanovsky V & Alon U Senescent cell turnover slows with age providing an explanation for the Gompertz law. Nat. Commun10, 5495 (2019). [PMC free article] [PubMed] [Google Scholar]
  274. Kang HT et al. Chemical screening identifies ATM as a target for alleviating senescence. Nat. Chem. Biol13, 616–623 (2017). [PubMed] [Google Scholar]
  275. Dimri GP et al. A biomarker that identifies senescent human cells in culture and in aging skin in vivo. Proc. Natl Acad. Sci. USA92, 9363–9367 (1995). [PMC free article] [PubMed] [Google Scholar]
  276. Hall BM et al. p16(Ink4a) and senescence-associated beta-galactosidase can be induced in macrophages as part of a reversible response to physiological stimuli. Aging9, 1867–1884 (2017). [PMC free article] [PubMed] [Google Scholar]
  277. Evangelou K et al. Robust, universal biomarker assay to detect senescent cells in biological specimens. Aging Cell16, 192–197 (2017). [PMC free article] [PubMed] [Google Scholar]
  278. Neurohr GE et al. excessive cell growth causes cytoplasm dilution and contributes to senescence. Cell176, 1083–1097 e1018 (2019). [PMC free article] [PubMed] [Google Scholar]
  279. Biran A et al. Quantitative identification of senescent cells in aging and disease. Aging Cell16, 661–671 (2017). [PMC free article] [PubMed] [Google Scholar]
  280. Witkiewicz AK, Knudsen KE, Dicker AP & Knudsen ES The meaning of p16(ink4a) expression in tumors: functional significance, clinical associations and future developments. Cell Cycle10, 2497–2503 (2011). [PMC free article] [PubMed] [Google Scholar]
  281. Herbig U, Ferreira M, Condel L, Carey D & Sedivy JM Cellular senescence in aging primates. Science311, 1257 (2006). [PubMed] [Google Scholar]
  282. Lozano-Torres B et al. The chemistry of senescence. Nat. Rev. Chem3, 426–441 (2019). [Google Scholar]
  283. Paez-Ribes M, González-Gualda E, Doherty GJ & Muñoz-Espín D Targeting senescent cells in translational medicine. EMBO Mol. Med11, e10234 (2019). [PMC free article] [PubMed] [Google Scholar]
  284. Wang Y et al. Real-time imaging of senescence in tumors with DNA damage. Sci. Rep9, 2102 (2019). [PMC free article] [PubMed] [Google Scholar]
  285. Doherty J & Baehrecke EH Life, death and autophagy. Nat. Cell Biol20, 1110–1117 (2018). [PubMed] [Google Scholar]
  286. Narita M et al. Spatial coupling of mTOR and autophagy augments secretory phenotypes. Science332, 966–970 (2011). [PMC free article] [PubMed] [Google Scholar]
  287. Young AR et al. Autophagy mediates the mitotic senescence transition. Genes Dev. 23, 798–803 (2009). [PMC free article] [PubMed] [Google Scholar]
  288. Dorr JR et al. Synthetic lethal metabolic targeting of cellular senescence in cancer therapy. Nature501, 421–425 (2013). [PubMed] [Google Scholar]
  289. Garcia-Prat L et al. Autophagy maintains stemness by preventing senescence. Nature529, 37–42 (2016). [PubMed] [Google Scholar]
  290. Tai H et al. Autophagy impairment with lysosomal and mitochondrial dysfunction is an important characteristic of oxidative stress-induced senescence. Autophagy13, 99–113 (2017). [PMC free article] [PubMed] [Google Scholar]

This is an open access article under the terms of the http://creativecommons.org/licenses/by/4.0/ License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

Credit also to PUBMED https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8344376/

Related Articles

We use cookies

We use cookies on our website. Some of them are essential for the operation of the site, while others help us to improve this site and the user experience (tracking cookies). You can decide for yourself whether you want to allow cookies or not. Please note that if you reject them, you may not be able to use all the functionalities of the site.