Skip to main content

Aging and cancer

Abstract

Aging and cancer exhibit apparent links that we will examine in this review. The null hypothesis that aging and cancer coincide because both are driven by time, irrespective of the precise causes, can be confronted with the idea that aging and cancer share common mechanistic grounds that are referred to as ‘hallmarks’. Indeed, several hallmarks of aging also contribute to carcinogenesis and tumor progression, but some of the molecular and cellular characteristics of aging may also reduce the probability of developing lethal cancer, perhaps explaining why very old age (> 90 years) is accompanied by a reduced incidence of neoplastic diseases. We will also discuss the possibility that the aging process itself causes cancer, meaning that the time-dependent degradation of cellular and supracellular functions that accompanies aging produces cancer as a byproduct or ‘age-associated disease’. Conversely, cancer and its treatment may erode health and drive the aging process, as this has dramatically been documented for cancer survivors diagnosed during childhood, adolescence, and young adulthood. We conclude that aging and cancer are connected by common superior causes including endogenous and lifestyle factors, as well as by a bidirectional crosstalk, that together render old age not only a risk factor of cancer but also an important parameter that must be considered for therapeutic decisions.

Introduction

Aging is the most important risk factor of malignant disease, the prevalence of which dramatically increases as adults age, reaching a peak around 85 or 90 years, when the incidence of new cancer diagnoses starts to decline and that of cardiovascular and other diseases ramps up [1, 2]. Aging is, to some degree, modulable, meaning that chronological age (measured in years) and biological age (measured by biological tests and clinical status) can be uncoupled from each other [3, 4]. A young biological age is linked to a reduced risk of malignant disease [5, 6]. For this reason, it may even be argued - in a polemic fashion - that aging is a modifiable risk factor of cancer. This speculation is apparently supported by epidemiological data indicating that lifestyle factors that slow the aging process - such as leanness, an equilibrated mostly plant-based diet, voluntary physical activity and the avoidance of environmental mutagens - also reduce the probability to develop malignant disease [7, 8]. This observation suggests - but does not prove - that aging and cancer share common causes that are influenced by lifestyle or, in a slightly different vision, that manifest aging precipitates the development of clinically detectable tumors that then develop as ‘age-related diseases’.

In this review, we will examine the mechanistic connections between aging and malignant disease (Fig. 1). We will first discuss arguments in favor of the null hypothesis (Fig. 1A), namely, that aging and cancer just coincide as we become older because both are time-dependent processes but do not necessarily share a common biological basis. This null hypothesis would be in line with the existence of childhood cancers and progeroid (i.e., aging-accelerating) syndromes that do not increase the likelihood to develop cancer. We will then examine the likely more broadly applicable hypothesis that aging and cancer have common mechanistic grounds, as supported by the idea that both these processes share molecular and cellular characteristics that have been referred to as ‘meta-hallmarks’ or ‘agonistic hallmarks’ (Fig. 1B). However, this hypothesis does not explain why very old age (> 90 years) is accompanied by a reduction of the incidence of cancers, perhaps because certain ‘antagonistic hallmarks’ of aging counteract carcinogenesis (Fig. 1C). There is also the possibility that aged tissues are more susceptible to the development and clinical manifestation of cancers that then develop as a consequence of biological aging (Fig. 1D). Conversely, cancer and its treatment with chemotherapy and radiotherapy can precipitate aging, reducing healthspan and lifespan, as this is well documented for childhood cancer survivors (CCSs) as well as for survivors of cancers treated during adolescence and young adulthood (Fig. 1E). Finally, we will discuss the importance to weigh therapeutic decisions as a function of the oncological patient’s biological age.

Fig. 1
figure 1

Potential relationship between aging and cancer. (A) Aging and cancer may lack a direct relationship and may rather be driven each independently by time (null hypothesis). (B) Agonistic drivers may cause aging and cancer in a time-dependent fashion. (C) Antagonistic drivers may favor aging while reducing the probability of carcinogenesis and tumor progression. (D) Aging tissues and organisms may be more prone for the development of cancers. (E) Cancer and its treatment may precipitate the deterioration of health and the aging process

The null hypothesis: no causal links between aging and cancer

Although most malignancies manifest in older adults (> 65 years) [1, 2], there are specific cancers that are diagnosed during childhood or adolescence without any accompanying signs of accelerated aging or the simultaneous development of other age-associated disorders such as cardiovascular and neurodegenerative diseases. Such early cancers are comparatively rare (~ 1 in 5000 of the under 20-year-old, accounting for just 1% of all cancer diagnoses) and mostly manifest as non-epithelial malignancies (e.g., leukemias, central nervous systems cancers and lymphomas), contrasting with older adults that preponderantly develop carcinomas, and appear uncoupled from the aging process [9, 10]. Conversely, it can be argued that such early-life cancers (as exemplified by germ cell tumors, hepatoblastomas, medulloblastomas, neuroblastomas, osteosarcomas, retinoblastomas, rhabdomyosarcomas, and Wilms tumors) have a peculiar molecular etiology, distinguishing them from the tumors developing in older adults. In addition, each of these malignancies peaks at a different age (1–2 years for neuroblastoma, 3–4 years for Wilms tumor, 5 years for rhabdomyosarcoma…) suggesting an association with specific developmental stages rather than cumulative alterations that occur during classical (i.e., age-associated) oncogenesis. In any case, it appears that a specific subgroup of cancers is uncoupled from aging.

Dissociation of aging and cancer is also observed for specific progeroid syndromes, i.e., genetically disorders resulting in premature and accelerated aging [11]. In sharp contrast with several progeroid syndromes caused by defects in DNA repair (e.g. Bloom syndrome, Werner syndrome and Xeroderma pigmentosa, XP), which are linked to the early manifestation of cancers that often occur in an organ-specific fashion (e.g. leukemia and lymphoma in Bloom syndrome; thyroid cancer, skin cancer, and sarcoma in Werner syndrome; ultraviolet light-induced skin cancer in XP) [12,13,14], other progeroid syndromes are not associated with any type of early carcinogenesis. Thus, trichothiodystrophy, which is caused by mutations in genes that are also mutated in XP (ERCC2, ERCC3) and do not only compromise DNA repair (as this occurs in XP) but also impair transcription (as this does not occur in XP), is not associated with malignant disease [15]. Similarly, Cockayne syndrome, which is caused by mutations affecting the transcription-coupled repair branch of the nucleotide excision repair pathway (ERCC6, ERCC8), photosensitizes the skin (as this applies to XP as well) but does not cause causer, likely because mutated cells are eliminated before they can transform to a malignant state [16].

The dissociation of accelerated aging phenotypes and cancer also applies to defects in lamin A/C, e.g. Hutchinson-Gilford progeria syndrome (HGPS) due to mutations in lamin A encoded by LMNA or its protease STE24 encoded by ZMPSTE24 [17]. As any other progeroid syndrome, HGPS causes segmental aging, i.e., an incomplete acquisition of aging phenotypes in only a few organ systems. Thus, HGPS patients develop some signs of aging (such as alopecia, wrinkled skin, osteoporosis, kidney failure, impaired vision and cardiovascular disease including atherosclerosis) but not others (such as cancer and neurodegeneration) during their infancy [18]. Similarly, Néstor–Guillermo progeria syndrome caused by BANF1 mutations is associated with an aged appearance and skeletal abnormalities but not others (such as cancer, diabetes, cardiovascular and neurodegenerative diseases) [19]. Thus, several progeroid syndromes do not lead to an increase in the incidence of cancers. However, given the extreme rarity of these syndromes (e.g., 1 in 10 to 20 million children for HGPS), it may be argued that they constitute again ‘exceptions that confirm the rule’. Moreover, the premature death caused by progeria (i.e., usually before 20 years in HGPS due to cardiovascular disease), might ‘hide’ their pro-oncogenic potential.

In conclusion, there is evidence that, in rare instances, aging phenotypes and cancer development can be uncoupled from each other. This applies to specific progeroid syndromes that, however, cause incomplete (segmental) aging, as well as to a specific array of cancers developing in children and adolescents that are molecularly different from tumors developing in older adults.

Common superior causes of aging and cancer

In contrast to the aforementioned exceptions, carcinomas, which constitute the most frequent category of cancers, as well as most glioblastomas, leukemias, lymphomas, melanomas and sarcomas, usually manifest at the age > 50 (in > 90% of all cases) and demonstrate a steady increase of incidence until the age of 85 years [1, 2]. Correlative evidence indicates that lifestyle factors that reduce biological aging also postpone or avoid the manifestation of cancer [20]. This applies to healthy lifestyles that increase organismal fitness including (i) a diverse, mostly plant-based diet based on natural ingredients (rather than highly processed foods, which are intrinsically toxic), avoiding overweight, obesity, hypovitaminoses, a deficit or surplus in oligoelements, as well as intestinal dysbiosis [21,22,23]; (ii) moderate or intense voluntary physical activity eluding excessive sedentarism, sarcopenia as well as osteoarthritis [24, 25]; (iii) avoidance of mutagenic toxins including excessive sun exposure, radiation, environmental poisons, air pollutants, tobacco and alcohol consumption [26, 27]; and (iv) psychosocial integration, which is often overlooked, yet essential for somatic health, in line with the fact that mental wellbeing and socioeconomic status are major determinants of healthspan, lifespan and the odds of cancer morbidity and mortality [28]. In accord with these observations, large epidemiological studies reveal that clinical factors for the most important age-associated ailments, i.e., cancer and cardiovascular disease, largely overlap [29, 30]. Of note, polygenic risk scores can predict the onset of both common cancers (such as mammary and prostate carcinoma) and cardiometabolic diseases [31].

The aforementioned associations between aging, cancer and cardiovascular disease suggest - but do not prove - that these conditions are dictated by common superior causes. What are then the hypothetical pathways that link such overarching mechanisms of aging and cancer? Such pathways can be tentatively identified among the ‘hallmarks’ of aging [3] and cancer [32], which do not only accompany the relevant processes but also accelerate them if they are experimentally or accidentally induced and, on the contrary, decelerate, halt or reverse aging as they simultaneously prevent carcinogenesis if they are attenuated by genetic or pharmacological manipulations [3, 32]. Several hallmarks of aging (i.e., genomic instability, epigenetic alterations, chronic inflammation and dysbiosis) are also described as hallmarks of cancer and hence constitute common ‘meta-hallmarks’ or ‘agonistic hallmarks’ [33] (Fig. 2).

Fig. 2
figure 2

Common mechanisms driving cancer and aging. Cancer and aging are characterized by common hallmarks: the chronic installation of inflammation, genomic instability, intestinal dysbiosis and alterations of the epigenome. ATM: ATM serine/threonine kinase; BLM: BLM RecQ like helicase; BRCA1/2: Breast cancer type 1/2 susceptibility protein; CRP: circRNA: circular RNA; C-reactive protein; ER: endoplasmic reticulum; ERCC8: Excision Repair Cross-Complementing group 8; FANCA/C/G: FA complementation group A/C/G; GABA: Gamma-aminobutyric acid; IFN-γ: Interferon-gamma; IL: interleukin; LAD: lamina-associated domains; lncRNA: long non-coding RNA; LPS: lipopolysaccharide; MDSC: myeloid-derived suppressor cells; miRNA: micro RNA; mRNA: messenger RNA; NF-κB: Nuclear factor kappa-light-chain-enhancer of activated B cells; NLRP3: NLR family pyrin domain containing 3; SCFA: short-chain fatty acids; TCR: T cell receptor; TH: helper T cell; TLR: Toll-like receptor; TNF-α: Tumor necrosis factor; TREG: regulatory T cell; WRN: WRN RecQ like helicase; 5-HT: 5-hydroxytryptamine (Serotonin)

Genomic instability

Mutations affecting chromosomal DNA occur spontaneously as well as in response to exogenous mutagens, resulting in a progressive, age-dependent accumulation of genomic alterations [34]. Next-generation sequencing of DNA extracted from circulating myeloid cells allows for the detection of clonal hematopoiesis of indetermined potential (CHIP). This alteration manifests with aging and constitutes a risk factor of blood cancers, including acute myeloid leukemia [35], as well as other seemingly unrelated diseases, such as atherosclerosis [36], liver fibrosis [37] and non-small cell lung cancer [38], likely due to pro-inflammatory effects. In recent years, it has been discovered that genomic instability affects all major organs, causing the generation of mosaics of cells (i.e., the juxtaposition of genetically non-identically cells within the same tissue), some of which tend to clonally expand because they acquire a proliferative advantage over normal, unmuted cells, hence outcompeting them [39, 40]. The resulting genetic heterogeneity may contribute to the time-dependent functional decline of aging tissues (for instance due to a final loss of stem cell features, replicative senescence, the secretion of pro-inflammatory factors) as well as the generation of ever-more mutated, pre-malignant and hence potentially oncogenic cells.

Epigenetic alterations

The structure of chromatin and patterns of gene expression are transmitted through epigenetic changes which result from a myriad of posttranslational modifications (most prominently methylation and acetylation) affecting DNA and histones (along with other mechanisms involving non-coding RNAs), as well as chromatin structure, that can be transmitted from mother cells to their daughter cells, hence contributing to the “identity” of differentiated cell types [41]. Throughout the aging process, such epigenetic changes are progressively lost, increasing the noise in the system, and contributing to a progressive loss of cellular identities that menaces the functional integrity of complex tissues and potentially enhances the risk of carcinogenesis coupled to an increase in tumor heterogeneity and phenotypic plasticity [42]. The most common (but still imperfect) technology to measure epigenetic shifts consists in bisulfite pyrosequencing to detect DNA methylation patterns that can be bioinformatically deconvoluted as “biological clocks” and be associated to the risks of developing specific diseases [43].

Chronic inflammation

Aging is associated with a failure to control inflammation in space and time (“inflamm-aging”) [44], and inflammation is also one of the hallmarks of cancer, likely acting through a combination of cell-autonomous effects (e.g., increased proliferation of cells leading to genomic and epigenomic instability) and non-cell-autonomous consequences (e.g., fibrosis, rarefaction of ECM components and local immunosuppression by myeloid-derived suppressor cells) [3, 45]. For this reason, inflammation has a dual role in both aging and cancer, implying that suppression of inflammation may have a multipronged impact on the development of a large spectrum of age-associated disorders that includes both malignant and non-malignant diseases.

Intestinal dysbiosis

The intestinal lumen is colonized by a diverse microecosystem composed by archaea, bacteria, fungi, parasites, phages and viruses that altogether influences gut health as well as bodywide homeostasis [46]. Contrasting with the healthy (eubiotic) state, gut dysbiosis is characterized by an increase in the abundance of harmful microbial species coupled to a relative decrease of useful microbes. Importantly, multiple non-malignant age-associated diseases are coupled with similar shifts in the gut microflora as are cancers located outside of the gastrointestinal tract [47]. Experiments showing that the microbiota from young mice, as well as specific health-associated bacterial strains (such as Akkermansia muciniphila), can enhance the lifespan of mice with progeria, suggest a causal implication of dysbiosis in aging [48]. Intriguingly, the transfer of such a health-associated microbiota or that of A. muciniphila also stimulate anticancer immunosurveillance [49] (and fecal microbial transfer from healthy patients to melanoma-bearing patients sensitizes to subsequent immunotherapy with antibodies targeting the PD-1/PD-L1 interaction) [50], suggesting communalities between the age-related loss of health and cancer. Thus, intestinal dysbiosis is considered another ‘meta-hallmark’ of aging and cancer.

In sum, it appears that some of the processes that cause aging also underly oncogenesis, as this is well documented for the accumulation of mutated cells in the aging organism, likely preparing the grounds for multi-step oncogenesis, as well as for the loss of epigenetically controlled cellular identities that may favor the acquisition of cancer stem cell characteristics. Chronic inflammation and dysbiosis also share similar etiologies and trajectories in the context of aging and cancer with the peculiarity that they can be targeted by specific treatments.

Possible causes of reduced cancer incidence in very old people

Nonagenarians (90–99 years), centenarians (100–109 years) and supercentenarians (> 110 years) progressively exhibit a relative decrease in the incidence of new cancer diagnoses as compared to the younger octogenarians (80–99 years) and septuagenarians (70–79 years) [1, 2], suggesting that some facets of the aging process may protect against the development and clinical manifestation of neoplasia. Indeed, the probability of a centenarian to die from cancer as opposed to other causes is only 4% [51]. Specific features of aging (i.e., telomere attrition and stem cell exhaustion) can suppress oncogenesis and hence act as ‘antagonistic’ hallmarks. Disabled macroautophagy and cellular senescence are two additional ‘ambivalent’ hallmarks of aging that mediate context-dependent oncosuppressive effects [33] (Fig. 3).

Fig. 3
figure 3

Mechanisms of aging that oppose cancer development. Part of the aging phenotype results in the blockade of mechanisms that typically sustain tumor development and growth. Telomere attrition, stem cell exhaustion, disabled macroautophagy and cellular senescence are increased in aging and have an antagonist role in cancer. ALT: Alternative lengthening of telomeres; ATG: autophagy-related genes; CCL2: chemokine C-C motif ligand 2; EP300: histone acetyltransferase p300; HH: hedgehog signaling pathway; hTERT: Telomerase reverse transcriptase; IFN: interferon; IL: interleukin; KLF4: Kruppel-like factor 4; MHC-I: major histocompatibility complex class I; mTORC1: mammalian target of rapamycin complex 1; Mφ: Macrophage; NK: natural killer cell; Notch: neurogenic locus notch homolog proteins signaling pathway; OCT4: octamer-binding transcription factor 4; p16: cyclin-dependent kinase inhibitor 2 A; p21: cyclin-dependent kinase inhibitor 1; RB1: Retinoblastoma protein; SASP: senescence-associated secretory phenotype; SOX2: sex determining region Y-box 2; TP53: tumor protein P53; WNT: Wnt signaling pathway

Telomere attrition

Telomeres at the extreme ends of chromosomes contain repeated sequences that must be maintained by the telomerase complex to avoid their progressive shortening during mitoses. Since telomerase subunits are typically lost during adulthood and aging in somatic cells, this mechanism limits replicative lifespan and potentially contributes to the aging process as a countdown mechanism [52]. Telomere attrition theoretically avoids carcinogenesis in aged tissues due to the induction of replicative senescence, and tumors must indeed re-activate telomerase expression (e.g., due to mutations in the promoter encoding the protein subunit TERT) [53], overexpress additional factors (such as the shelterin compound TPP1) that cooperate with telomerase in telomere maintenance [54], or activate mechanisms for alternative lengthening of telomers to strive [55].

Stem cell exhaustion

Stem cell exhaustion compromises tissue repair in aging [56, 57]. Although this has negative effects on the capacity of tissues to regenerate upon injury, stem cell exhaustion may also prevent oncogenesis by opposing phenotypic plasticity and hence reduce the probability of malignant transformation in aged tissues [33]. In other words, stem cell exhaustion can abort the first steps of oncogenesis, which relies on the formation cancer stem cells. Indeed, malignant transformation implies a failure of normal terminal differentiation by cells that rather undergo de-differentiation, manifest a differentiation block or exhibit transdifferentiation [32]. Some of these pathways related to phenotypic plasticity (such as signals transmitted via Wnt/β-catenin, NF-κB, Hedgehog) are explored in clinical trials [58] and Smoothened (Smo) antagonists can be targeted for the treatment of locally advanced and metastatic basal cell carcinoma [59], underscoring the practical relevance of these findings.

Disabled macroautophagy

Aging is associated to a progressive inhibition of macroautophagy (and other types of autophagy, including chaperone-mediated autophagy and mitophagy), progressively compromising cellular fitness due to the accumulation of waste material including dysfunctional organelles and micronuclei [60, 61]. Disabled macroautophagy may also compromise the fitness of cancer cells, reducing their metabolic fitness, proliferative potential, resistance to therapeutic agents, as well as their capacity to subvert anticancer immune responses [33]. That, said, macroautophagy may also constitute a tumor-suppressive mechanism because it contributes to the maintenance of genomic stability, favors oncogene-induced senescence, mitigates procarcinogenic inflammation, contributes to ferroptotic cell death [62] and favors immunosurveillance [33, 63]. Hence, it appears that macroautophagy plays a context-dependent role, either favoring or inhibiting oncogenesis and tumor progression.

Cellular senescence

Senescent cells exhibiting accumulate in aging tissues, and it has been postulated that their close-to-irreversible cell cycle arrest would constitute a barrier against malignant transformation [33]. Accordingly, the induction of senescence in malignant cells may constitute a therapeutic goal, especially since senescent cancer cells appear to be particularly immunogenic, hence eliciting T cell responses via the upregulation of the antigen-presenting machinery in response to interferon-g [64, 65]. In addition, senescent tumor cells appear particularly susceptible to natural killer (NK) cell-mediated lysis [66, 67]. However, senescence may be reversible in specific cases, a phenomenon that might contribute to tumor cell dormancy [68, 69]. Moreover, senescence can result in local immunosuppression due to upregulation of the two PD-1 ligands PD-L1 and/or PD-L2 on malignant cells [70, 71], as well as in the secretion of pro-inflammatory and immunosuppressive factors exemplified by interleukins 6 and 8 [72]. This latter phenomenon, which is dubbed as senescence-associated secretory phenotype (SASP), explains the long-range effects of cellular senescence [73]. When senescence affects tumor-infiltrating leukocytes, it may subvert anticancer immune responses (see below), hence contributing to tumor progression. Furthermore, senescence affecting stromal cells (such as hepatic stellate cells in the context of hepatocellular carcinoma) may precipitate oncogenesis [74]. For this reason, senescence mediates context-dependent anti- and pro-carcinogenic effects.

In sum, several among the hallmarks of aging may reduce the generation or fitness of (pre-)malignant cells, likely explaining why the oldest elderly exhibit a reduced cancer-specific mortality. That said, although such aging-associated tumor suppressive effects may have a significant impact on cancer development in very old persons, they fall short from reducing cancer incidence to the levels found before 20 years of age [1, 2].

Cancer as a complication of aging

The aforementioned considerations suggest that cancers steadily increase their frequency in the aged organism until the plateau reached at 85–90 years is attained, because aging and oncogenesis are caused by shared mechanisms (and simultaneously other aging-driving processes fail to avoid carcinogenesis). However, it can also be speculated that age-associated changes in tissue quality with fibrosis and alterations of the extracellular matrix (ECM), systemic and local inflammation, as well as failure of immunosurveillance favor carcinogenesis and tumor progression [75, 76]. Hence aging itself (rather than its underlying causes) would support the clinical manifestation and progression of cancers as a secondary complication of aging (Fig. 4).

Fig. 4
figure 4

Reciprocal induction of aging and cancer

The aged organism is particularly propitious for the development of malignancies due to alterations in the extracellular matrix (ECM) and the installation of a favorable immune context (inflammation and immunosenescence). Conversely, after their curative treatment cancer survivors face long-term toxicities including accelerated aging. Indeed, in the long run, they have higher probabilities of cancer relapse as well as increased risk of developing a plethora of age-related pathologies. CD: Cluster of differentiation; CSF: colony-stimulating factors; CXCL1: chemokine (C-X-C motif) ligand 1; FRP1/2: secreted frizzled-related proteins 1/2; HAPLN1: hyaluronan and proteoglycan link protein 1; ICAM1: intercellular adhesion molecule 1; IL: interleukin; irAEs: immune-related adverse events; MDSC: myeloid-derived suppressor cells; MMPs: matrix metalloproteinases; NO: nitric oxide; PA: protease associated domain proteins; PDGF-C: platelet-derived growth factor C; ROS: reactive oxygen species; TCR: T cell receptor; VEGF: vascular endothelial growth factor

Alterations of the extracellular matrix

Aging is associated to the development of fibrosis due to the excessive deposition of ECM components such as collagen in the ECM in several internal organs. This property may explain why aging is coupled to an increased propensity of breast cancers and melanomas to generate metastases in the lung. Indeed, in preclinical experiments, fibrosis of the lung causes the reversal of dormancy of cancer cells via the fibroblast-mediated secretion of platelet-derived growth factor (PDGF)-C (in the case of estrogen receptor-positive breast cancer) or that of WNT antagonist, sFRP1 (in the case of melanoma) [77, 78]. Reportedly, aged dermal fibroblasts also secrete high levels of another WNT antagonist, sFRP2, which can drive angiogenesis in melanomas, their metastasis, as well as their resistance to targeted therapy with the BRAF inhibitor vemurafenib [79]. In addition, age-associated disruption of the collagen I network in the ECM of the dermis may reduce mechanical constraints that prevent the development of basal cell carcinoma [80].

An age-related decrease in the secreted ECM polysaccharide hyaluronic acid, especially in its high-molecular mass variant, may causally contribute to aging and oncogenesis, as demonstrated by the fact that transgenic mice overexpressing naked mole-rat hyaluronic acid synthase 2 gene exhibit an increase in cancer-free healthspan and longevity [81]. This age-associated decrease in hyaluronic acid, as well as that of the proteoglycan link protein hyaluronan and proteoglycan link protein 1 (HAPLN1), may induce an aging-associated increase in ICAM1 in endothelial cells [82]. ICAM1 overexpression causes phosphorylation and internalization of VE-cadherin, resulting in blood vessel permeabilization, potentially explaining why old age is associated with poor melanoma outcome. Indeed, blocking ICAM1 with suitable antibodies reduces tumor size and distant metastasis in older mice with melanoma [82].

Inflammation

Inflammaging [44] can drive the senescence of cancer-associated fibroblasts that secrete factors enhancing peritoneal dissemination of gastric cancer [83]. Gliosis, a state of central nervous system inflammation coupled to the expansion of glial cells (such as microglia and astrocytes causing microgliosis and astrogliosis, respectively, during early and late responses to injury) promotes metastasis of lymphoma to the brain due to the upregulation of the chemokine CCL19, locally retaining tumor cells [84]. As compared to plasma from young controls, plasma from aged individuals contains higher levels of methylmalonic acid, a byproduct of propionate catabolism and a biomarker of vitamin B12 deficiency [85]. B12 deficiency may favor inflammation indirectly through a failure in tissue repair [86]. Of note, methylmalonic acid favors epithelial-mesenchymal transition of cancer cells through the upregulation of TGFB2 and consequent upregulation of the transcription factor SOX4 [85]. In addition, methylmalonic acid has pro-inflammatory and pro-aging properties [87]. These examples illustrate how age-associated inflammation favors tumor progression.

Failing immunosurveillance

Aging of the immune system (immunosenescence) occurs in the elderly, thus compromising anticancer immune response that may avoid carcinogenesis, reduce tumor progression, and decisively contribute to the success of most if not all treatment modalities in the oncological armamentarium, including chemotherapy, radiotherapy, immunotherapy and targeted therapy [88, 89]. Immunosenescence may directly affect T cells, reducing their effector function by down-regulating the costimulatory markers CD28 and CD27 and upregulating the terminal differentiation marker CD57 [90, 91]. In addition, senescent macrophages accumulate in tissues such as the lung, facilitating KRAS-induced non-small cell lung cancers, likely due to direct trophic effects on malignant cells, as well as due to the suppression of T cell-mediated immunosurveillance. Accordingly, the elimination of senescent macrophages reduces tumor progression [92, 93]. Failing immunosurveillance may also contribute to aging due to the incapacity of the immune system to clear senescent cells that accumulate in various tissues. Logically, attempts are underway to stimulate immune responses against such senescent cells, for instance by engineering chimeric antigen receptor (CAR) T cells that recognize antigens associated with cellular senescence [94, 95].

In conclusion, the aging organism appears particularly susceptible to the development and progression of malignant tumors through a variety of mechanisms. Aging tissues may constitute a particularly appropriate ‘soil’ for tumors to seed and invade.

Aging as a consequence of cancer and its treatment

Invasive cancers break tissue barriers, cause chronic inflammation, suppress immune responses, and mobilize ever more resources from the body, ultimately eroding bodywide health at multiple levels [96]. Moreover, even when successful, their treatment with DNA-damaging chemotherapeutics and radiotherapy has long-lasting effects on the organism that may manifest with a delay of several decades in cancer survivors cured during childhood, adolescence, or young adulthood. These long-term consequences give rise to a premature aging phenotype coupled to the early manifestation of a large panel of age-associated pathologies that include, but are not limited to, the manifestation of other (‘subsequent’ or ‘second’) cancers, small adult height, prediabetes, cardiovascular disease, chronic kidney disease, dementia, musculoskeletal decline with osteoporosis and sarcopenia, as well as tissue fibrosis. Ultimately, this results in frailty and early mortality (Fig. 4). These long-term complications of early-life cancer treatments have been described in some detail thanks to the constitution of specific registries such as the St. Jude Lifetime Cohort [97, 98], the US-centered Childhood Cancer Survivor study [98, 99], and the EUROCARE-6 study [100]. Although not as obvious as observed in childhood cancer, the additive burden of previous cancer in terms of chronic pathologies and premature mortality can be calculated in the adult population. Using data from the UK biobank, the health data from over 240,000 cancer survivors was compared to that of 500,000 adults with no history of cancer after matching by age, sex, and Index of Multiple Deprivation. Late morbidities attributable to cancer included hematological, pulmonary, Immune and renal dysfunctions, and depended on the type, doses and combination of used therapies [101]. Logically, attempts are underway to palliate these undesired side effects by more appropriate treatments reducing long-term toxicity, screening programs that identify patients at risk of developing specific diseases, as well as by post-therapeutic lifestyle interventions.

Avoidance of long-term toxicities of anticancer treatments

There is clear evidence that the severity of the age- and disease-accelerating effects of early-life cancer therapies have diminished over time likely due to several factors including, but not limited to, the reduction of cumulative chemotherapy doses, the replacement of some DNA-damaging agents by other cytotoxicants, and the avoidance of certain interventions, such as cranial irradiation of children with leukemia or glioma; or mediastinal irradiation of patients with Hodgkin lymphoma [102, 103]. Retrospective analyses identifying risk-enhancing practices and biomarkers may help to reduce treatment-induced long-term toxicities in prospective studies. Thus, telomere length in circulating lymphocytes is reduced in CCSs, correlating with the manifestation of a variety of non-neoplastic chronic health conditions [104]. Similarly, the measurement of various signs of biological aging (two physiology-based algorithms; four distinct DNA methylation clocks, and a single-time-point DNA methylation blood test) revealed that CCSs from the St. Jude Liftetime cohort aged more quickly (by ~ 5% in average) than community controls, in particular when they received hematopoietic cell transplants and vinca alkaloid chemotherapy [105]. Although these quantitative tests are predictive of mortality [105], it remains to be determined whether such biomarkers may guide the development of less toxic cancer cures.

An additional strategy consists in the use of co-medications that can reduce anticancer drug toxicities. For example, co-treatment with the iron chelator dexrazoxane has been successfully used to mitigate the long-term side effects of anthracyclines at the level of serious cardiovascular outcomes (cardiomyopathy, ischemic heart disease, and stroke) in CCSs [94, 95]. Moreover, in a randomized Phase II trial, low-dose tamoxifen has been shown to reduce radiological and biological risk factors of breast cancer in patients having received chest radiation ≥ 12 Gy by the age of 40 [106]. Anthracycline-induced premature aging can be prevented in mice by a chemical-genetic system that allows for the elimination of senescent cells [107]. Hence, senolytics, which are drugs that kill senescent cells, can be used to combat the long-term cardiotoxicity of doxorubicin in a preclinical model [108]. Future will tell whether such an approach can also be used to mitigate therapy-induced senescence in cancer patients as well.

Biomarker-guided screening programs

The risk of subsequent (secondary) cancers can be calculated based on polygenic risk scores derived from general population and genome-wide association studies [109]. Moreover, this risk is influenced by the type of treatment (radiotherapy and specific chemotherapeutic agents) and their cumulative doses [110]. More than cumulative interactions between genetic risk and radiotherapy have been described for specific cancers such as basal cell carcinoma and breast or thyroid cancers [109]. The use of doxorubicin beyond a threshold (≥ 200 mg.m-2) is linked to an enhanced risk of subsequent female breast cancer [110], exemplifying how distinct therapeutic interventions are linked to particular cancer risks that may instigate a reinforcement of early detection campaigns. Among CCSs, hearing loss is associated with the use of cisplatin, carboplatin and cranial or facial radiation > 32 Gy [111]. The risk of cardiac failure is determined by cumulative anthracycline doses and the location of radiotherapyn e.g., targeting the mediastinum causing irradiation of the heart [112, 113], while the risk of severe obesity in CCSs is influenced by genetic risk scores [114]. Thus, particular features of early-life cancer therapy may be combined with polygenic risk scores to guide specific screening programs for the detection and interception of specific manifestations of premature aging in CCSs.

Lifestyle interventions

Retrospective studies indicate that premature aging of CCSs is reduced by enhanced uptake of dark green vegetables and nuts/seeds, but enhanced by that of refined grain [115]. In contrast it appears that physical activity has no significant impact on the probability of CCSs to develop subsequent cancers [116, 117]. However, physical activity in adult CCSs has been shown to correlate with reduced neurocognitive problems at the levels of emotion regulation, memory, organization and task efficiency [118], as well as with reduced mortality [116]. In addition, psychosocial stress, sleep perturbations, smoking, alcohol consumption and substance use may contribute to accelerated ageing in CCSs [119] in the same way as they deteriorate health in cancer-free individuals [28]. These findings suggest that lifestyle factors that favor health in the general population may also be useful for maintaining the fitness of CCSs.

In sum, survivors of early-life cancer exhibit accelerated aging with the precocious manifestation of age-associated diseases, as well as an elevated risk of frailty and premature death. Attempts are underway to reduce these risks. Thus, secondary prevention in childhood cancer survivors is constantly ameliorated following specific guidelines, such as the Children’s Oncology Group Long-Term Follow-Up Guidelines for Survivors of Childhood, Adolescent, and Young Adult Cancers in North America [120], the Pan-European Network for Care of Survivors after Childhood and Adolescent Cancer Guidelines Group [121], as well as the International Guideline Harmonization Group for Late Effects of Childhood Cancer [122]. Beyond these risk reduction programs, efforts are ongoing to implement lifestyle interventions that reduce accelerated aging in cancer survivors. It will be interesting to learn whether drugs that are currently evaluated for their potential antiaging effects in clinical trials [4] can be advantageously used in cancer survivors as well.

Impact of aging on the therapeutic management of cancer

The classification of cancer is still mostly based on location (organs) rather than on molecular subtypes. When classified by location or histology, the prognosis of each cancer type changes with age [123]. For example, breast cancers tend to be particularly aggressive if they manifest before 40 years of age [124], while Hodgkin lymphoma diagnosed after 45 has a dismal prognosis compared to cases diagnosed in adolescence or early adulthood [125]. Similarly, the efficacy of treatments regimens changes with age. For instance, oxaliplatin fails to confer any benefit for the adjuvant treatment of poor-prognosis colorectal cancer after the age of 70 [125], but immunotherapy against melanoma or lung cancer is equally efficient at a young and an old age [126, 127]. Considering that adjuvant or neoadjuvant chemotherapy of breast cancer patients leads to a 10% reduction of exercise capacity, measured by oxygen uptake during peak exercise (VO2peak), and that normal aging is accompanied by a 10% reduction of VO2peak per decade, the age acceleration induced by therapeutic interventions on adult patients is certainly problematic [128]. For this reason, generic recommendations such as the avoidance of chemotherapy and a preference for radiotherapy for the management of older cancer patients have been proposed [129]. However, this idea collides with the fact that the suppression of chemotherapy in older breast cancer patients is associated with an elevated risk of relapse [130].

The majority of cancers manifest in older adults (> 65 years), often in the context of advanced biological age (with respect to chronological age) and one or several comorbidities. This contrasts with the fact that most clinical trials are performed in younger, relatively fit individuals, because they usually exclude persons > 70 years with major comorbidities and reduced performance status [131, 132], meaning that FDA/EMA-approved treatments are often not adapted to the average ‘real world’ cancer patient. Indeed, older adults diagnosed with cancer may exhibit more side effects and reduced drug tolerability than patients enrolled in clinical trials. For this reason, it is necessary to carefully weight therapeutic decisions to avoid the over-treatment or under-treatment of older patients. Over-treatment consists in surgical procedures or the administration of excessive doses (of drugs or irradiation) or cycles of treatments, resulting in a reduction of the quality of life without therapeutic benefit, as this often occurs near the end of life in older patients [133]. Under-treatment consists in the exclusion of older patients from viable therapeutic options based on the mere consideration of their chronological age, without taking into account their biological fitness [134, 135].

To adapt cancer therapies to each cancer patient in a personalized fashion, recommendations have been formulated by American Society for Clinical Oncology [136], Federal Drug Administration [137], the Cancer and Aging Research Group [138], and the International Society of Geriatric Oncology Priorities Initiative [139]. This involves comprehensive geriatric assessment (GA) of patients before therapeutic decisions are made, ideally in the context of a medical team involving both geriatricians and oncologists. GA should include the combined evaluation of physical performance, functional status, comorbidities, polypharmacy, cognition, nutrition, social support, and psychological status [136]. GA then allows to classify patients into fit, vulnerable, and frail. Fit patients can be oriented towards standard of care, vulnerable individuals towards interventions that reduce geriatric conditions as they undergo adapted treatments (e.g., with reduced doses and number of cycles or giving preference to radiotherapy over chemotherapy), and frail persons towards palliative care [140] (Fig. 5). Randomized clinical studies demonstrated that GA can reduce serious toxic effects from cancer treatment [141]. Beyond GA, it is possible to measure biological parameters indicating health deterioration among older cancer patients such as the levels of circulating C-reactive protein, a parameter of systemic inflammation, to predict other parameters such as cognitive decline [142]. Indeed, it has been proposed to measure multiple parameters indicative of inflammation, cell senescence, telomere shortening, and epigenetic changes that may inform on the biological resilience of older cancer patients and then influence treatment decisions [143].

Fig. 5
figure 5

Practical management of geriatric patients after cancer diagnosis

Flow chart for the adaptation of the general cancer clinical management guidelines to the specific needs of the geriatric population

In synthesis, clinical oncology is confronted with the challenge of adapting treatments to a heterogeneous population of mostly elderly patients that differ in their biological and medical conditions. In this context, a major challenge is to transcend the idea that the extension of overall survival (quantity of life) constitutes the sole desirable endpoint and hence to consider the importance of quality-of-life as well [135].

Conclusions

In this review, we have outlined some of the overarching principles governing the relationship between aging and cancer. Aging is strongly linked to cancer at three levels, namely, (i) because aging and oncogenesis share common mechanisms, (ii) because aging tissues favor tumor progression, and (iii) because tumor therapies undermine health and cause premature aging. Exceptions to these rules are constituted by (i) pediatric cancers that preferentially manifest during infancy rather than adulthood, (ii) the existence of progeroid syndromes without malignancies, and (iii) the fact that the oldest elderly exhibit a reduced incidence of new diagnoses of, and death from, cancer.

Worldwide estimations indicate that 1.6 billion individuals will be over 65 in 2050, implying a major surge in the number of age-related diseases including cancer. In this context, it will be important to decipher the precise mechanisms that link old age to the manifestation and progression of neoplasia and to develop broadly implementable strategies for the prevention, early detection and interception of malignant disease, hence avoiding the diagnosis of cancer at an advanced stage, when treatments become poorly tolerable, expensive, and mostly futile. Hence, investments in public and private research dealing with aging and cancer should be a priority for the future. Such investments will not only provide a molecular comprehension of the crosstalk between aging and malignancy, but will also lead to the identification of actionable targets for prophylactic or early-interceptive interventions on both processes.

It is reasonable to postulate that lifestyle interventions coupled to public policies designed to reduce exposure to industrial, nutritional, and environmental pollutants and to improve the economic and psychosocial status of the aging population, will allow to extend healthspan and to delay or avoid the manifestation of neoplastic disease. In this context, different countries have organized their pension and health systems, anti-pollutant strategies, as well as their focus on preventive versus curative medical interventions, in rather distinct ways. It will be a challenge for future investigation to perform carefully controlled inter-country comparisons so that the outcome of such policies can be accurately interpreted and improved. By applying policies that are successful in one country to others and by performing sophisticated performance measurements, it should be possible to perform large-scale multidisciplinary studies that will optimize a sustainable society that efficiently supports the prevention and interception of old age-associated cancer.

Data availability

No datasets were generated or analysed during the current study.

Abbreviations

CAR:

chimeric antigen receptor

CCS:

childhood cancer survivor

CHIP:

clonal hematopoiesis of indetermined potential

ECM:

extracellular matri

GA:

geriatric assessment

HGPS:

Hutchinson-Gilford progeria syndrome

HAPLN2:

hyaluronan and proteoglycan link protein 1

NK:

natural killer

PDGF:

platelet-derived growth factor

XP:

xeroderma pigmentosum

References

  1. Siegel RL, Giaquinto AN, Jemal A. Cancer statistics, 2024. CA Cancer J Clin. 2024;74(1):12–49.

    Article  PubMed  Google Scholar 

  2. Global age-sex. -specific mortality, life expectancy, and population estimates in 204 countries and territories and 811 subnational locations, 1950–2021, and the impact of the COVID-19 pandemic: a comprehensive demographic analysis for the global burden of Disease Study 2021. Lancet. 2024.

  3. López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G. Hallmarks of aging: an expanding universe. Cell. 2023;186(2):243–78.

    Article  PubMed  Google Scholar 

  4. Guarente L, Sinclair DA, Kroemer G. Human trials exploring anti-aging medicines. Cell Metab. 2024;36(2):354–76.

    Article  CAS  PubMed  Google Scholar 

  5. Zheng Y, Joyce BT, Colicino E, Liu L, Zhang W, Dai Q, et al. Blood epigenetic age may predict cancer and mortality. EBioMedicine. 2016;5:68–73.

    Google Scholar 

  6. Jia Q, Chen C, Xu A, Wang S, He X, Shen G, et al. A biological age model based on physical examination data to predict mortality in a Chinese population. iScience. 2024;27(3):108891.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Matthews CE, Moore SC, Arem H, Cook MB, Trabert B, Håkansson N, et al. Amount and intensity of leisure-time physical activity and lower cancer risk. J Clin Oncol. 2020;38(7):686–97.

    Article  PubMed  Google Scholar 

  8. Karavasiloglou N, Thompson AS, Pestoni G, Knuppel A, Papier K, Cassidy A, et al. Adherence to the EAT-lancet reference diet is associated with a reduced risk of incident cancer and all-cause mortality in UK adults. One Earth. 2023;6(12):1726–34.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Ward ZJ, Yeh JM, Bhakta N, Frazier AL, Atun R. Estimating the total incidence of global childhood cancer: a simulation-based analysis. Lancet Oncol. 2019;20(4):483–93.

    Article  PubMed  Google Scholar 

  10. Johnston WT, Erdmann F, Newton R, Steliarova-Foucher E, Schüz J, Roman E. Childhood cancer: estimating regional and global incidence. Cancer Epidemiol. 2021;71:101662. Pt B).

    Article  CAS  PubMed  Google Scholar 

  11. Carrero D, Soria-Valles C, López-Otín C. Hallmarks of progeroid syndromes: lessons from mice and reprogrammed cells. Dis Model Mech. 2016;9(7):719–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Lebel M, Monnat RJ. Jr. Werner syndrome (WRN) gene variants and their association with altered function and age-associated diseases. Ageing Res Rev. 2018;41:82–97.

    Article  CAS  PubMed  Google Scholar 

  13. Rizza ERH, DiGiovanna JJ, Khan SG, Tamura D, Jeskey JD, Kraemer KH. Xeroderma pigmentosum: a model for human premature aging. J Invest Dermatol. 2021;141(4s):976–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ababou M. Bloom syndrome and the underlying causes of genetic instability. Mol Genet Metab. 2021;133(1):35–48.

    Article  CAS  PubMed  Google Scholar 

  15. Lombardi A, Arseni L, Carriero R, Compe E, Botta E, Ferri D, et al. Reduced levels of prostaglandin I(2) synthase: a distinctive feature of the cancer-free trichothiodystrophy. Proc Natl Acad Sci U S A. 2021;118:26.

    Article  Google Scholar 

  16. Reid-Bayliss KS, Arron ST, Loeb LA, Bezrookove V, Cleaver JE. Why cockayne syndrome patients do not get cancer despite their DNA repair deficiency. Proc Natl Acad Sci U S A. 2016;113(36):10151–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Gordon LB, Rothman FG, López-Otín C, Misteli T. Progeria: a paradigm for translational medicine. Cell. 2014;156(3):400–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Batista NJ, Desai SG, Perez AM, Finkelstein A, Radigan R, Singh M et al. The molecular and cellular basis of Hutchinson-Gilford progeria syndrome and potential treatments. Genes (Basel). 2023;14(3).

  19. Cabanillas R, Cadiñanos J, Villameytide JA, Pérez M, Longo J, Richard JM, et al. Néstor-Guillermo progeria syndrome: a novel premature aging condition with early onset and chronic development caused by BANF1 mutations. Am J Med Genet A. 2011;155a(11):2617–25.

    Article  PubMed  Google Scholar 

  20. Kroemer G, McQuade JL, Merad M, André F, Zitvogel L. Bodywide ecological interventions on cancer. Nat Med. 2023;29(1):59–74.

    Article  CAS  PubMed  Google Scholar 

  21. Kliemann N, Rauber F, Bertazzi Levy R, Viallon V, Vamos EP, Cordova R, et al. Food processing and cancer risk in Europe: results from the prospective EPIC cohort study. Lancet Planet Health. 2023;7(3):e219–32.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Montégut L, de Cabo R, Zitvogel L, Kroemer G. Science-Driven nutritional interventions for the prevention and treatment of cancer. Cancer Discov. 2022;12(10):2258–79.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Tang D, Kroemer G, Kang R. Targeting cuproplasia and cuproptosis in cancer. Nat Rev Clin Oncol. 2024.

  24. Qiu Y, Fernández-García B, Lehmann HI, Li G, Kroemer G, López-Otín C, et al. Exercise sustains the hallmarks of health. J Sport Health Sci. 2023;12(1):8–35.

    Article  PubMed  Google Scholar 

  25. Shreves AH, Small SR, Travis RC, Matthews CE, Doherty A. Dose-response of accelerometer-measured physical activity, step count, and cancer risk in the UK Biobank: a prospective cohort analysis. Lancet. 2023;402(Suppl 1):S83.

    Article  PubMed  Google Scholar 

  26. Anthony KM, Collins JM, Love SM, Stewart JD, Buchheit SF, Gondalia R, et al. Radon exposure, clonal hematopoiesis, and stroke susceptibility in the women’s Health Initiative. Neurology. 2024;102(2):e208055.

    Article  CAS  PubMed  Google Scholar 

  27. Hill W, Lim EL, Weeden CE, Lee C, Augustine M, Chen K, et al. Lung adenocarcinoma promotion by air pollutants. Nature. 2023;616(7955):159–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. López-Otín C, Kroemer G. The missing hallmark of health: psychosocial adaptation. Cell Stress. 2024;8:21–50.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Abdellatif M, Rainer PP, Sedej S, Kroemer G. Hallmarks of cardiovascular ageing. Nat Rev Cardiol. 2023;20(11):754–77.

    Article  PubMed  Google Scholar 

  30. Lau ES, Paniagua SM, Liu E, Jovani M, Li SX, Takvorian K, et al. Cardiovascular risk factors are associated with future cancer. JACC CardioOncol. 2021;3(1):48–58.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Mars N, Koskela JT, Ripatti P, Kiiskinen TTJ, Havulinna AS, Lindbohm JV, et al. Polygenic and clinical risk scores and their impact on age at onset and prediction of cardiometabolic diseases and common cancers. Nat Med. 2020;26(4):549–57.

    Article  CAS  PubMed  Google Scholar 

  32. Hanahan D. Hallmarks of cancer: new dimensions. Cancer Discov. 2022;12(1):31–46.

    Article  CAS  PubMed  Google Scholar 

  33. López-Otín C, Pietrocola F, Roiz-Valle D, Galluzzi L, Kroemer G. Meta-hallmarks of aging and cancer. Cell Metab. 2023;35(1):12–35.

    Article  PubMed  Google Scholar 

  34. Niedernhofer LJ, Gurkar AU, Wang Y, Vijg J, Hoeijmakers JHJ, Robbins PD. Nuclear genomic instability and aging. Annu Rev Biochem. 2018;87:295–322.

    Article  CAS  PubMed  Google Scholar 

  35. Sperling AS, Gibson CJ, Ebert BL. The genetics of myelodysplastic syndrome: from clonal haematopoiesis to secondary leukaemia. Nat Rev Cancer. 2017;17(1):5–19.

    Article  CAS  PubMed  Google Scholar 

  36. Jaiswal S, Libby P. Clonal haematopoiesis: connecting ageing and inflammation in cardiovascular disease. Nat Rev Cardiol. 2020;17(3):137–44.

    Article  PubMed  Google Scholar 

  37. Wong WJ, Emdin C, Bick AG, Zekavat SM, Niroula A, Pirruccello JP, et al. Clonal haematopoiesis and risk of chronic liver disease. Nature. 2023;616(7958):747–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Tian R, Wiley B, Liu J, Zong X, Truong B, Zhao S, et al. Clonal hematopoiesis and risk of incident lung cancer. J Clin Oncol. 2023;41(7):1423–33.

    Article  CAS  PubMed  Google Scholar 

  39. Kakiuchi N, Ogawa S. Clonal expansion in non-cancer tissues. Nat Rev Cancer. 2021;21(4):239–56.

    Article  CAS  PubMed  Google Scholar 

  40. Yokoyama A, Kakiuchi N, Yoshizato T, Nannya Y, Suzuki H, Takeuchi Y, et al. Age-related remodelling of oesophageal epithelia by mutated cancer drivers. Nature. 2019;565(7739):312–7.

    Article  CAS  PubMed  Google Scholar 

  41. Davalos V, Esteller M. Cancer epigenetics in clinical practice. CA Cancer J Clin. 2023;73(4):376–424.

    Article  PubMed  Google Scholar 

  42. Feinberg AP, Levchenko A. Epigenetics as a mediator of plasticity in cancer. Science. 2023;379(6632):eaaw3835.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Haghani A, Li CZ, Robeck TR, Zhang J, Lu AT, Ablaeva J, et al. DNA methylation networks underlying mammalian traits. Science. 2023;381(6658):eabq5693.

    Article  CAS  PubMed  Google Scholar 

  44. Franceschi C, Bonafè M, Valensin S, Olivieri F, De Luca M, Ottaviani E, et al. Inflamm-aging. An evolutionary perspective on immunosenescence. Ann N Y Acad Sci. 2000;908:244–54.

    Article  CAS  PubMed  Google Scholar 

  45. Swanton C, Bernard E, Abbosh C, André F, Auwerx J, Balmain A, et al. Embracing cancer complexity: Hallmarks of systemic disease. Cell. 2024;187(7):1589–616.

    Article  CAS  PubMed  Google Scholar 

  46. Routy B, Gopalakrishnan V, Daillere R, Zitvogel L, Wargo JA, Kroemer G. The gut microbiota influences anticancer immunosurveillance and general health. Nat Rev Clin Oncol. 2018;15(6):382–96.

    Article  CAS  PubMed  Google Scholar 

  47. Thomas AM, Fidelle M, Routy B, Kroemer G, Wargo JA, Segata N, et al. Gut OncoMicrobiome signatures (GOMS) as next-generation biomarkers for cancer immunotherapy. Nat Rev Clin Oncol. 2023;20(9):583–603.

    Article  PubMed  Google Scholar 

  48. Barcena C, Valdes-Mas R, Mayoral P, Garabaya C, Durand S, Rodriguez F, et al. Healthspan and lifespan extension by fecal microbiota transplantation into progeroid mice. Nat Med. 2019;25(8):1234–42.

    Article  CAS  PubMed  Google Scholar 

  49. Routy B, Le Chatelier E, Derosa L, Duong CPM, Alou MT, Daillere R, et al. Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science. 2018;359(6371):91–7.

    Article  CAS  PubMed  Google Scholar 

  50. Routy B, Lenehan JG, Miller WH Jr., Jamal R, Messaoudene M, Daisley BA, et al. Fecal microbiota transplantation plus anti-PD-1 immunotherapy in advanced melanoma: a phase I trial. Nat Med. 2023;29(8):2121–32.

    Article  CAS  PubMed  Google Scholar 

  51. Joseph SC, Delcastilo E, Loukas M, Osiro S. Common cancers in centenarians. Med Sci Monit. 2014;20:18–23.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Roake CM, Artandi SE. Regulation of human telomerase in homeostasis and disease. Nat Rev Mol Cell Biol. 2020;21(7):384–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Tornesello ML, Cerasuolo A, Starita N, Amiranda S, Bonelli P, Tuccillo FM, et al. Reactivation of telomerase reverse transcriptase expression in cancer: the role of TERT promoter mutations. Front Cell Dev Biol. 2023;11:1286683.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Chun-On P, Hinchie AM, Beale HC, Gil Silva AA, Rush E, Sander C, et al. TPP1 promoter mutations cooperate with TERT promoter mutations to lengthen telomeres in melanoma. Science. 2022;378(6620):664–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Gao J, Pickett HA. Targeting telomeres: advances in telomere maintenance mechanism-specific cancer therapies. Nat Rev Cancer. 2022;22(9):515–32.

    Article  CAS  PubMed  Google Scholar 

  56. Brown RL. Stem cell exhaustion and atherosclerosis. J Anti Aging Med. 2003;6(3):279. discussion 80.

    Article  PubMed  Google Scholar 

  57. Lopez-Otin C, Blasco MA, Partridge L, Serrano M, Kroemer G. Hallmarks of aging: an expanding universe. Cell. 2023;186(2):243–78.

    Article  CAS  PubMed  Google Scholar 

  58. Yang L, Shi P, Zhao G, Xu J, Peng W, Zhang J, et al. Targeting cancer stem cell pathways for cancer therapy. Signal Transduct Target Ther. 2020;5(1):8.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Ruat M, Hoch L, Faure H, Rognan D. Targeting of smoothened for therapeutic gain. Trends Pharmacol Sci. 2014;35(5):237–46.

    Article  CAS  PubMed  Google Scholar 

  60. Uddin MN, Nishio N, Ito S, Suzuki H, Isobe K. Autophagic activity in thymus and liver during aging. Age (Dordr). 2012;34(1):75–85.

    Article  PubMed  Google Scholar 

  61. Aman Y, Schmauck-Medina T, Hansen M, Morimoto RI, Simon AK, Bjedov I, et al. Autophagy in healthy aging and disease. Nat Aging. 2021;1(8):634–50.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Chen X, Tsvetkov AS, Shen H-M, Isidoro C, Ktistakis NT, Linkermann A et al. International consensus guidelines for the definition, detection, and interpretation of autophagy-dependent ferroptosis. Autophagy. 2024:1–34.

  63. Vodnala SK, Eil R, Kishton RJ, Sukumar M, Yamamoto TN, Ha N-H, et al. T cell stemness and dysfunction in tumors are triggered by a common mechanism. Volume 363. New York, NY: Science; 2019. p. eaau0135. 6434.

    Google Scholar 

  64. Chen H-A, Ho Y-J, Mezzadra R, Adrover JM, Smolkin R, Zhu C, et al. Senescence rewires microenvironment sensing to facilitate antitumor immunity. Cancer Discov. 2023;13(2):432–53.

    Article  CAS  PubMed  Google Scholar 

  65. Marin I, Boix O, Garcia-Garijo A, Sirois I, Caballe A, Zarzuela E, et al. Cellular senescence is immunogenic and promotes antitumor immunity. Cancer Discov. 2023;13(2):410–31.

    Article  CAS  PubMed  Google Scholar 

  66. Colucci M, Zumerle S, Bressan S, Gianfanti F, Troiani M, Valdata A, et al. Retinoic acid receptor activation reprograms senescence response and enhances anti-tumor activity of natural killer cells. Cancer Cell. 2024;42(4):646–e619.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Ruscetti M, Leibold J, Bott MJ, Fennell M, Kulick A, Salgado NR, et al. NK cell-mediated cytotoxicity contributes to tumor control by a cytostatic drug combination. Volume 362. New York, NY: Science; 2018. pp. 1416–22. 6421.

    Google Scholar 

  68. Igelmann S, Lessard F, Uchenunu O, Bouchard J, Fernandez-Ruiz A, Rowell MC, et al. A hydride transfer complex reprograms NAD metabolism and bypasses senescence. Mol Cell. 2021;81(18):3848–e6519.

    Article  CAS  PubMed  Google Scholar 

  69. Saleh T, Tyutyunyk-Massey L, Gewirtz DA. Tumor cell escape from therapy-induced senescence as a model of disease recurrence after dormancy. Cancer Res. 2019;79(6):1044–6.

    Article  CAS  PubMed  Google Scholar 

  70. Chaib S, Lopez-Dominguez JA, Lalinde-Gutierrez M, Prats N, Marin I, Boix O, et al. The efficacy of chemotherapy is limited by intratumoral senescent cells expressing PD-L2. Nat Cancer. 2024;5(3):448–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Shahbandi A, Chiu FY, Ungerleider NA, Kvadas R, Mheidly Z, Sun MJS, et al. Breast cancer cells survive chemotherapy by activating targetable immune-modulatory programs characterized by PD-L1 or CD80. Nat Cancer. 2022;3(12):1513–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Faget DV, Ren Q, Stewart SA. Unmasking senescence: context-dependent effects of SASP in cancer. Nat Rev Cancer. 2019;19(8):439–53.

    Article  CAS  PubMed  Google Scholar 

  73. Wang B, Han J, Elisseeff JH, Demaria M. The senescence-associated secretory phenotype and its physiological and pathological implications. Nat Rev Mol Cell Biol. 2024.

  74. Yamagishi R, Kamachi F, Nakamura M, Yamazaki S, Kamiya T, Takasugi M, et al. Gasdermin D-mediated release of IL-33 from senescent hepatic stellate cells promotes obesity-associated hepatocellular carcinoma. Sci Immunol. 2022;7(72):eabl7209.

    Article  CAS  PubMed  Google Scholar 

  75. Jassim A, Rahrmann EP, Simons BD, Gilbertson RJ. Cancers make their own luck: theories of cancer origins. Nat Rev Cancer. 2023;23(10):710–24.

    Article  CAS  PubMed  Google Scholar 

  76. Yuan S, Almagro J, Fuchs E. Beyond genetics: driving cancer with the tumour microenvironment behind the wheel. Nat Rev Cancer. 2024;24(4):274–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Fane ME, Chhabra Y, Alicea GM, Maranto DA, Douglass SM, Webster MR, et al. Stromal changes in the aged lung induce an emergence from melanoma dormancy. Nature. 2022;606(7913):396–405.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Turrell FK, Orha R, Guppy NJ, Gillespie A, Guelbert M, Starling C, et al. Age-associated microenvironmental changes highlight the role of PDGF-C in ER(+) breast cancer metastatic relapse. Nat Cancer. 2023;4(4):468–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Kaur A, Webster MR, Marchbank K, Behera R, Ndoye A, Kugel CH 3, et al. sFRP2 in the aged microenvironment drives melanoma metastasis and therapy resistance. Nature. 2016;532(7598):250–4.

  80. Bansaccal N, Vieugue P, Sarate R, Song Y, Minguijon E, Miroshnikova YA, et al. The extracellular matrix dictates regional competence for tumour initiation. Nature. 2023;623(7988):828–35.

    Article  CAS  PubMed  Google Scholar 

  81. Zhang Z, Tian X, Lu JY, Boit K, Ablaeva J, Zakusilo FT, et al. Increased hyaluronan by naked mole-rat Has2 improves healthspan in mice. Nature. 2023;621(7977):196–205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Marino-Bravante GE, Carey AE, Huser L, Dixit A, Wang V, Kaur A, et al. Age-dependent loss of HAPLN1 erodes vascular integrity via indirect upregulation of endothelial ICAM1 in melanoma. Nat Aging. 2024;4(3):350–63.

    Article  CAS  PubMed  Google Scholar 

  83. Yasuda T, Koiwa M, Yonemura A, Miyake K, Kariya R, Kubota S, et al. Inflammation-driven senescence-associated secretory phenotype in cancer-associated fibroblasts enhances peritoneal dissemination. Cell Rep. 2021;34(8):108779.

    Article  CAS  PubMed  Google Scholar 

  84. O’Connor T, Zhou X, Kosla J, Adili A, Garcia Beccaria M, Kotsiliti E, et al. Age-related gliosis promotes central nervous system lymphoma through CCL19-mediated tumor cell retention. Cancer Cell. 2019;36(3):250–67. e9.

    Article  PubMed  Google Scholar 

  85. Gomes AP, Ilter D, Low V, Endress JE, Fernandez-Garcia J, Rosenzweig A, et al. Age-induced accumulation of methylmalonic acid promotes tumour progression. Nature. 2020;585(7824):283–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Kovatcheva M, Melendez E, Chondronasiou D, Pietrocola F, Bernad R, Caballe A, et al. Vitamin B(12) is a limiting factor for induced cellular plasticity and tissue repair. Nat Metab. 2023;5(11):1911–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Tejero J, Lazure F, Gomes AP. Methylmalonic acid in aging and disease. Trends Endocrinol Metab. 2024;35(3):188–200.

    Article  CAS  PubMed  Google Scholar 

  88. Kroemer G, Chan TA, Eggermont AMM, Galluzzi L. Immunosurveillance in clinical cancer management. CA Cancer J Clin. 2024;74(2):187–202.

    Article  PubMed  Google Scholar 

  89. Zitvogel L, Tesniere A, Kroemer G. Cancer despite immunosurveillance: immunoselection and immunosubversion. Nat Rev Immunol. 2006;6(10):715–27.

    Article  CAS  PubMed  Google Scholar 

  90. Vicente R, Mausset-Bonnefont AL, Jorgensen C, Louis-Plence P, Brondello JM. Cellular senescence impact on immune cell fate and function. Aging Cell. 2016;15(3):400–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Huang M, Wang Y, Fang L, Liu C, Feng F, Liu L, et al. T cell senescence: a new perspective on immunotherapy in lung cancer. Front Immunol. 2024;15:1338680.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Haston S, Gonzalez-Gualda E, Morsli S, Ge J, Reen V, Calderwood A, et al. Clearance of senescent macrophages ameliorates tumorigenesis in KRAS-driven lung cancer. Cancer Cell. 2023;41(7):1242–60. e6.

    Article  CAS  PubMed  Google Scholar 

  93. Prieto LI, Sturmlechner I, Graves SI, Zhang C, Goplen NP, Yi ES, et al. Senescent alveolar macrophages promote early-stage lung tumorigenesis. Cancer Cell. 2023;41(7):1261–75. e6.

    Article  CAS  PubMed  Google Scholar 

  94. Amor C, Feucht J, Leibold J, Ho YJ, Zhu C, Alonso-Curbelo D, et al. Senolytic CAR T cells reverse senescence-associated pathologies. Nature. 2020;583(7814):127–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Baker DJ, Arany Z, Baur JA, Epstein JA, June CH. CAR T therapy beyond cancer: the evolution of a living drug. Nature. 2023;619(7971):707–15.

    Article  CAS  PubMed  Google Scholar 

  96. Lopez-Otin C, Kroemer G. Hallmarks of health. Cell. 2021;184(7):1929–39.

    Article  CAS  PubMed  Google Scholar 

  97. Dixon SB, Wang F, Lu L, Wilson CL, Green DM, Merchant TE, et al. Prediabetes and associated risk of cardiovascular events and chronic kidney disease among adult survivors of childhood cancer in the St Jude lifetime cohort. J Clin Oncol. 2024;42(9):1031–43.

    Article  CAS  PubMed  Google Scholar 

  98. Ehrhardt MJ, Krull KR, Bhakta N, Liu Q, Yasui Y, Robison LL, et al. Improving quality and quantity of life for childhood cancer survivors globally in the twenty-first century. Nat Rev Clin Oncol. 2023;20(10):678–96.

    Article  PubMed  Google Scholar 

  99. Dieffenbach BV, Murphy AJ, Liu Q, Ramsey DC, Geiger EJ, Diller LR, et al. Cumulative burden of late, major surgical intervention in survivors of childhood cancer: a report from the childhood cancer survivor study (CCSS) cohort. Lancet Oncol. 2023;24(6):691–700.

    Article  PubMed  Google Scholar 

  100. Botta L, Gatta G, Capocaccia R, Stiller C, Canete A, Dal Maso L, et al. Long-term survival and cure fraction estimates for childhood cancer in Europe (EUROCARE-6): results from a population-based study. Lancet Oncol. 2022;23(12):1525–36.

    Article  PubMed  Google Scholar 

  101. Chang WH, Neal RD, Forster MD, Lai AG. Cumulative burden of 144 conditions, critical care hospitalisation and premature mortality across 26 adult cancers. Nat Commun. 2023;14(1):1484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. de Blank PMK, Lange KR, Xing M, Mirzaei Salehabadi S, Srivastava D, Brinkman TM et al. Temporal changes in treatment and late mortality and morbidity in adult survivors of childhood glioma: a report from the Childhood Cancer Survivor Study. Nat Cancer. 2024.

  103. Dixon SB, Chen Y, Yasui Y, Pui CH, Hunger SP, Silverman LB, et al. Reduced morbidity and mortality in survivors of childhood acute lymphoblastic leukemia: a report from the childhood cancer survivor study. J Clin Oncol. 2020;38(29):3418–29.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Song N, Li Z, Qin N, Howell CR, Wilson CL, Easton J, et al. Shortened leukocyte telomere length associates with an increased prevalence of chronic health conditions among survivors of childhood cancer: a report from the St. Jude lifetime cohort. Clin Cancer Res. 2020;26(10):2362–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Guida JL, Hyun G, Belsky DW, Armstrong GT, Ehrhardt MJ, Hudson MM et al. Associations of seven measures of biological age acceleration with frailty and all-cause mortality among adult survivors of childhood cancer in the St. Jude lifetime cohort. Nat Cancer. 2024.

  106. Bhatia S, Palomares MR, Hageman L, Chen Y, Landier W, Smith K, et al. A randomized phase IIb study of low-dose tamoxifen in chest-irradiated cancer survivors at risk for breast cancer. Clin Cancer Res. 2021;27(4):967–74.

    Article  CAS  PubMed  Google Scholar 

  107. Demaria M, O’Leary MN, Chang J, Shao L, Liu S, Alimirah F, et al. Cellular senescence promotes adverse effects of chemotherapy and cancer relapse. Cancer Discov. 2017;7(2):165–76.

    Article  CAS  PubMed  Google Scholar 

  108. Lerida-Viso A, Estepa-Fernandez A, Morella-Aucejo A, Lozano-Torres B, Alfonso M, Blandez JF, et al. Pharmacological senolysis reduces doxorubicin-induced cardiotoxicity and improves cardiac function in mice. Pharmacol Res. 2022;183:106356.

    Article  CAS  PubMed  Google Scholar 

  109. Gibson TM, Karyadi DM, Hartley SW, Arnold MA, Berrington de Gonzalez A, Conces MR, et al. Polygenic risk scores, radiation treatment exposures and subsequent cancer risk in childhood cancer survivors. Nat Med. 2024;30(3):690–8.

    Article  CAS  PubMed  Google Scholar 

  110. Wang Y, Ronckers CM, van Leeuwen FE, Moskowitz CS, Leisenring W, Armstrong GT, et al. Subsequent female breast cancer risk associated with anthracycline chemotherapy for childhood cancer. Nat Med. 2023;29(9):2268–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Beyea JA, Lau C, Cooke B, Hall S, Nathan PC, Gupta S. Long-term incidence and predictors of significant hearing loss requiring hearing assistive devices among childhood cancer survivors: a population-based study. J Clin Oncol. 2020;38(23):2639–46.

    Article  PubMed  Google Scholar 

  112. de Baat EC, Feijen EAM, Reulen RC, Allodji RS, Bagnasco F, Bardi E, et al. Risk factors for heart failure among Pan-european Childhood Cancer survivors: a PanCareSurFup and ProCardio cohort and nested case-control study. J Clin Oncol. 2023;41(1):96–106.

    Article  PubMed  Google Scholar 

  113. de Vries S, Haaksma ML, Jóźwiak K, Schaapveld M, Hodgson DC, Lugtenburg PJ, et al. Development and validation of risk prediction models for coronary heart disease and heart failure after treatment for Hodgkin Lymphoma. J Clin Oncol. 2023;41(1):86–95.

    Article  PubMed  Google Scholar 

  114. Sapkota Y, Qiu W, Dixon SB, Wilson CL, Wang Z, Zhang J, et al. Genetic risk score enhances the risk prediction of severe obesity in adult survivors of childhood cancer. Nat Med. 2022;28(8):1590–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Wang M, Lan T, Williams AM, Ehrhardt MJ, Lanctot JQ, Jiang S et al. Plant foods intake and risk of premature aging in adult survivors of childhood cancer in the St Jude lifetime cohort (SJLIFE). J Clin Oncol. 2024:JCO2301260.

  116. Scott JM, Li N, Liu Q, Yasui Y, Leisenring W, Nathan PC, et al. Association of exercise with mortality in adult survivors of childhood cancer. JAMA Oncol. 2018;4(10):1352–8.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Morales JS, Valenzuela PL, Velázquez-Díaz D, Castillo-García A, Jiménez-Pavón D, Lucia A et al. Exercise and childhood cancer-A historical review. Cancers (Basel). 2021;14(1).

  118. Barlow-Krelina E, Chen Y, Yasui Y, Till C, Gibson TM, Ness KK, et al. Consistent physical activity and future neurocognitive problems in adult survivors of childhood cancers: a report from the childhood cancer survivor study. J Clin Oncol. 2020;38(18):2041–52.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Carroll JE, Bower JE, Ganz PA. Cancer-related accelerated ageing and biobehavioural modifiers: a framework for research and clinical care. Nat Rev Clin Oncol. 2022;19(3):173–87.

    Article  CAS  PubMed  Google Scholar 

  120. Landier W, Bhatia S, Eshelman DA, Forte KJ, Sweeney T, Hester AL, et al. Development of risk-based guidelines for pediatric cancer survivors: the children’s oncology group long-term follow-up guidelines from the children’s oncology group late effects committee and nursing discipline. J Clin Oncol. 2004;22(24):4979–90.

    Article  PubMed  Google Scholar 

  121. van Kalsbeek RJ, van der Pal HJH, Kremer LCM, Bardi E, Brown MC, Effeney R, et al. European PanCareFollowUp recommendations for surveillance of late effects of childhood, adolescent, and young adult cancer. Eur J Cancer. 2021;154:316–28.

    Article  PubMed  Google Scholar 

  122. Kremer LC, Mulder RL, Oeffinger KC, Bhatia S, Landier W, Levitt G, et al. A worldwide collaboration to harmonize guidelines for the long-term follow-up of childhood and young adult cancer survivors: a report from the international late effects of childhood cancer guideline harmonization group. Pediatr Blood Cancer. 2013;60(4):543–9.

    Article  PubMed  Google Scholar 

  123. Shah Y, Verma A, Marderstein AR, White J, Bhinder B, Garcia Medina JS, et al. Pan-cancer analysis reveals molecular patterns associated with age. Cell Rep. 2021;37(10):110100.

    Article  CAS  PubMed  Google Scholar 

  124. Kim HJ, Kim S, Freedman RA, Partridge AH. The impact of young age at diagnosis (age < 40 years) on prognosis varies by breast cancer subtype: a U.S. SEER database analysis. Breast. 2022;61:77–83.

    Article  PubMed  Google Scholar 

  125. Rose A, Grajales-Cruz A, Lim A, Todd A, Bello C, Shah B, et al. Classical Hodgkin Lymphoma: clinicopathologic features, prognostic factors, and outcomes from a 28-Year single institutional experience. Clin Lymphoma Myeloma Leuk. 2021;21(2):132–8.

    Article  PubMed  Google Scholar 

  126. Nebhan CA, Cortellini A, Ma W, Ganta T, Song H, Ye F, et al. Clinical outcomes and toxic effects of single-agent immune checkpoint inhibitors among patients aged 80 years or older with cancer: a Multicenter International Cohort Study. JAMA Oncol. 2021;7(12):1856–61.

    Article  PubMed  Google Scholar 

  127. Presley CJ, Gomes F, Burd CE, Kanesvaran R, Wong ML. Immunotherapy in older adults with cancer. J Clin Oncol. 2021;39(19):2115–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Shafqat S, Arana Chicas E, Shafqat A, Hashmi SK. The Achilles’ heel of cancer survivors: fundamentals of accelerated cellular senescence. J Clin Invest. 2022;132(13).

  129. Amini A, Morris L, Ludmir EB, Movsas B, Jagsi R, VanderWalde NA. Radiation therapy in older adults with cancer: a critical modality in geriatric oncology. J Clin Oncol. 2022;40(16):1806–11.

    Article  CAS  PubMed  Google Scholar 

  130. Crozier JA, Pezzi TA, Hodge C, Janeva S, Lesnikoski BA, Samiian L, et al. Addition of chemotherapy to local therapy in women aged 70 years or older with triple-negative breast cancer: a propensity-matched analysis. Lancet Oncol. 2020;21(12):1611–9.

    Article  PubMed  Google Scholar 

  131. Le-Rademacher J, Mohile S, Unger J, Hudson MF, Foster J, Lichtman S, et al. Trial design considerations to increase older adult accrual to national cancer institute clinical trials. J Natl Cancer Inst Monogr. 2022;2022(60):135–41.

    Article  PubMed  PubMed Central  Google Scholar 

  132. Sedrak MS, Freedman RA, Cohen HJ, Muss HB, Jatoi A, Klepin HD, et al. Older adult participation in cancer clinical trials: a systematic review of barriers and interventions. CA Cancer J Clin. 2021;71(1):78–92.

    Article  PubMed  Google Scholar 

  133. Fang P, Jagsi R, He W, Lei X, Campbell EG, Giordano SH, et al. Rising and falling trends in the use of chemotherapy and targeted therapy near the end of life in older patients with cancer. J Clin Oncol. 2019;37(20):1721–31.

    Article  PubMed  PubMed Central  Google Scholar 

  134. DuMontier C, Loh KP, Bain PA, Silliman RA, Hshieh T, Abel GA, et al. Defining undertreatment and overtreatment in older adults with cancer: a scoping literature review. J Clin Oncol. 2020;38(22):2558–69.

    Article  PubMed  PubMed Central  Google Scholar 

  135. DuMontier C, Loh KP, Soto-Perez-de-Celis E, Dale W. Decision making in older adults with cancer. J Clin Oncol. 2021;39(19):2164–74.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Dale W, Klepin HD, Williams GR, Alibhai SMH, Bergerot C, Brintzenhofeszoc K, et al. Practical assessment and management of vulnerabilities in older patients receiving systemic cancer therapy: ASCO guideline update. J Clin Oncol. 2023;41(26):4293–312.

    Article  PubMed  Google Scholar 

  137. F.D.A. Guidance for Industry - Inclusion of Older Adults in Cancer Clinical Trials. FDA-2019-D-55722022. p. https://www.fda.gov/regulatory-information/search-fda-guidance-documents/inclusion-older-adults-cancer-clinical-trials

  138. Hurria A, Akiba C, Kim J, Mitani D, Loscalzo M, Katheria V, et al. Reliability, validity, and feasibility of a computer-based geriatric assessment for older adults with cancer. J Oncol Pract. 2016;12(12):e1025–34.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Extermann M, Brain E, Canin B, Cherian MN, Cheung KL, de Glas N, et al. Priorities for the global advancement of care for older adults with cancer: an update of the international society of geriatric oncology priorities initiative. Lancet Oncol. 2021;22(1):e29–36.

    Article  PubMed  Google Scholar 

  140. Ioffe D, Dotan E. Evidence-based care of older adults with metastatic colorectal cancer: insights from landmark clinical trials. J Clin Oncol. 2023;41(34):5228–36.

    Article  PubMed  Google Scholar 

  141. Mohile SG, Mohamed MR, Xu H, Culakova E, Loh KP, Magnuson A, et al. Evaluation of geriatric assessment and management on the toxic effects of cancer treatment (GAP70+): a cluster-randomised study. Lancet. 2021;398(10314):1894–904.

    Article  PubMed  PubMed Central  Google Scholar 

  142. Carroll JE, Nakamura ZM, Small BJ, Zhou X, Cohen HJ, Ahles TA, et al. Elevated C-reactive protein and subsequent patient-reported cognitive problems in older breast Cancer survivors: the thinking and living with cancer study. J Clin Oncol. 2023;41(2):295–306.

    Article  CAS  PubMed  Google Scholar 

  143. Sedrak MS, Gilmore NJ, Carroll JE, Muss HB, Cohen HJ, Dale W. Measuring biologic resilience in older cancer survivors. J Clin Oncol. 2021;39(19):2079–89.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Figures were created with BioRender.com.

Funding

GK is supported by the Ligue contre le Cancer (équipe labellisée); Agence National de la Recherche (ANR) – Projets blancs; Association pour la recherche sur le cancer (ARC); Cancéropôle Ile-de-France; Fondation pour la Recherche Médicale (FRM); a donation by Elior; European Joint Programme on Rare Diseases (EJPRD) Wilsonmed; European Research Council Advanced Investigator Award (ERC-2021-ADG, Grant No. 101052444; project acronym: ICD-Cancer, project title: Immunogenic cell death (ICD) in the cancer-immune dialogue), The ERA4 Health Cardinoff Grant Ener-LIGHT, European Union Horizon 2020 research and innovation programmes Oncobiome (grant agreement number: 825410, Project Acronym: ONCOBIOME, Project title: Gut OncoMicrobiome Signatures [GOMS] associated with cancer incidence, prognosis and prediction of treatment response, Prevalung (grant agreement number 101095604, Project Acronym: PREVALUNG EU, project title: Biomarkers affecting the transition from cardiovascular disease to lung cancer: towards stratified interception), Neutrocure (grant agreement number 861878: Project Acronym: Neutrocure; project title: Development of “smart” amplifiers of reactive oxygen species specific to aberrant polymorphonuclear neutrophils for treatment of inflammatory and autoimmune diseases, cancer and myeloablation); National support managed by the Agence Nationale de la Recherche under the France 2030 programme (reference number 21-ESRE-0028, ESR/Equipex + Onco-Pheno-Screen); Hevolution Network on Senescence in Aging; Institut National du Cancer (INCa); Institut Universitaire de France; LabEx Immuno-Oncology ANR-18-IDEX-0001; a Cancer Research ASPIRE Award from the Mark Foundation; PAIR-Obésité INCa_1873, the RHUs Immunolife and LUCA-pi (both dedicated to France Relance 2030); Seerave Foundation; SIRIC Cancer Research and Personalized Medicine (CARPEM). This study contributes to the IdEx Université de Paris Cité ANR-18-IDEX-0001. Views and opinions expressed are those of the author(s) only and do not necessarily reflect those of the European Union, the European Research Council or any other granting authority. Neither the European Union nor any other granting authority can be held responsible for them.

Author information

Authors and Affiliations

Authors

Contributions

G.K. and L.M. performed literature searches, wrote the text and designed figures. C.L.-O. edited the text, provided intellectual input and suggested figures. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Guido Kroemer.

Ethics declarations

Ethical approval

This is a review article, meaning that no ethical approval is needed.

Competing interests

GK has been holding research contracts with Daiichi Sankyo, Eleor, Kaleido, Lytix Pharma, PharmaMar, Osasuna Therapeutics, Samsara Therapeutics, Sanofi, Sutro, Tollys, and Vascage. GK is on the Board of Directors of the Bristol Myers Squibb Foundation France. GK is a scientific co-founder of everImmune, Osasuna Therapeutics, Samsara Therapeutics and Therafast Bio. GK is in the scientific advisory boards of Hevolution, Institut Servier, Longevity Vision Funds and Rejuveron Life Sciences. GK is the inventor of patents covering therapeutic targeting of aging, cancer, cystic fibrosis and metabolic disorders. GK’s wife, Laurence Zitvogel, has held research contracts with Glaxo Smyth Kline, Incyte, Lytix, Kaleido, Innovate Pharma, Daiichi Sankyo, Pilege, Merus, Transgene, 9 m, Tusk and Roche, was on the on the Board of Directors of Transgene, is a cofounder of everImmune, and holds patents covering the treatment of cancer and the therapeutic manipulation of the microbiota. GK’s brother, Romano Kroemer, was an employee of Sanofi and now consults for Boehringer-Ingelheim. The funders had no role in the design of the study; in the writing of the manuscript, or in the decision to publish the results.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Montégut, L., López-Otín, C. & Kroemer, G. Aging and cancer. Mol Cancer 23, 106 (2024). https://doi.org/10.1186/s12943-024-02020-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-024-02020-z

Keywords