Skip to main content

REVIEW article

Front. Cardiovasc. Med., 20 October 2021
Sec. Cardio-Oncology
This article is part of the Research Topic Cardio-Oncology and Reverse Cardio-Oncology: the Manifold Interconnections Between Heart Failure and Cancer View all 8 articles

Senescence-Associated Secretory Phenotype as a Hinge Between Cardiovascular Diseases and Cancer

  • 1Center for Cardiovascular Sciences, Houston Methodist Research Institute, Houston, TX, United States
  • 2Department of Cardiology, The University of Texas MD Anderson Cancer Center, Houston, TX, United States
  • 3Department of Pediatrics Research, The University of Texas MD Anderson Cancer Center, Houston, TX, United States

Overlapping risks for cancer and cardiovascular diseases (CVD), the two leading causes of mortality worldwide, suggest a shared biology between these diseases. The role of senescence in the development of cancer and CVD has been established. However, its role as the intersection between these diseases remains unclear. Senescence was originally characterized by an irreversible cell cycle arrest after a high number of divisions, namely replicative senescence (RS). However, it is becoming clear that senescence can also be instigated by cellular stress, so-called stress-induced premature senescence (SIPS). Telomere shortening is a hallmark of RS. The contribution of telomere DNA damage and subsequent DNA damage response/repair to SIPS has also been suggested. Although cellular senescence can mediate cell cycle arrest, senescent cells can also remain metabolically active and secrete cytokines, chemokines, growth factors, and reactive oxygen species (ROS), so-called senescence-associated secretory phenotype (SASP). The involvement of SASP in both cancer and CVD has been established. In patients with cancer or CVD, SASP is induced by various stressors including cancer treatments, pro-inflammatory cytokines, and ROS. Therefore, SASP can be the intersection between cancer and CVD. Importantly, the conventional concept of senescence as the mediator of cell cycle arrest has been challenged, as it was recently reported that chemotherapy-induced senescence can reprogram senescent cancer cells to acquire “stemness” (SAS: senescence-associated stemness). SAS allows senescent cancer cells to escape cell cycle arrest with strongly enhanced clonogenic growth capacity. SAS supports senescent cells to promote both cancer and CVD, particularly in highly stressful conditions such as cancer treatments, myocardial infarction, and heart failure. As therapeutic advances have increased overlapping risk factors for cancer and CVD, to further understand their interaction may provide better prevention, earlier detection, and safer treatment. Thus, it is critical to study the mechanisms by which these senescence pathways (SAS/SASP) are induced and regulated in both cancer and CVD.

Introduction

The health and physiological state of humans or any animal is governed by tissue homeostasis which is significantly controlled by physiological and environmental signals (1, 2). In response to potential damage signals, cellular machinery activates the damage response system to reverse damage to the cells through various mechanisms, as have been reviewed extensively elsewhere (35). However, when the damage is irreparable, the cells often undergo a programmed cell death, or apoptosis, in combination with tissue necrosis (5). Distinct from these two extreme phenomena is another cell fate called “senescence” (6, 7). The concept of cellular senescence (from the Latin word “senex” meaning “old”) was first introduced by Hayflick and Moorhead in 1961 when they observed that in cell culture, human diploid fibroblasts were irreversibly arrested after serial passaging (8). This limited replicative/proliferative capacity was named replicative senescence (RS) (8, 9). In addition to RS, cellular senescence can be induced by both extra- and intra-cellular stimuli including genotoxic agents, stress, mitochondrial dysfunction, nutrient deficit, radiation, and oncogene activation, so-called stress-induced premature senescence (SIPS) (Figure 1). In this review, we will focus on how senescence, especially SIPS, contribute to the progression of cancer and cardiovascular diseases (CVD), which may be the key to understanding the interconnection between them.

FIGURE 1
www.frontiersin.org

Figure 1. Various forms of senescence.

Telomeric DNA Damage, But Not Telomere Shortening, Induces Sips

Following cell divisions, telomere length is shortened to a critical level at which cells can no longer replicate and enter RS (913). Therefore, telomere shortening has a key role in RS. Alternatively, SIPS is different from RS in terms of molecular mechanisms and time frame. SIPS is induced by oxidative stress or DNA damaging agents in a relatively short period of time (usually 3–10 days) with or without significant telomere shortening (14). Both genomic and telomeric DNA damages can induce SIPS (15). However, most genomic DNA damages can be repaired by the DNA damage response (DDR) mechanisms within 24 h after stress (16), while telomeric DNA damages persist for months (17). Therefore, telomeric DNA damage-induced SIPS may explain the late effects triggered by various stressors including cancer treatments, as we will describe in the next sections. Importantly, telomeric DNA damages are occurred despite the shortening of telomere length and the expression of telomerase enzyme (17, 18). The dispensable role of telomere shortening in the development of senescence was also confirmed by the study showing that in human cancer cells, the very long telomeres were found to be more sensitive to ionizing radiation (IR) (19). Parrinello et al. reported that 20% oxygen density induces SIPS in mouse embryonic fibroblast without telomere shortening (14). Magalhães et al. reported that Ultraviolet B or hydrogen peroxide (H2O2) induces senescence markers, p21(WAF-1) and p16(INK4a), and increases senescence associated β-galactosidase (SA-β-gal) staining without provoking telomere shortening in telomerase immortalized human foreskin fibroblast, hTERT-BJ1 (20). An analysis of the telomere in the small airway epithelial cells from the lungs of patients suffering from Chronic Obstructive Pulmonary Disease showed that p16(INK4a) was highly expressed in those cells while the telomere length was not significantly shorter (21). Overall, these data suggest that telomere shortening is dispensable for SIPS, and that stress-induced telomeric DNA damages and the subsequent DDR, but not telomere shortening, is important for SIPS (20).

Senescence-Associated Secretory Phenotype (SASP) Can be Induced by Both RS and Sips

Senescent cells produce and release a variety of factors, including inflammatory cytokines (such as interleukin (IL)-1,−1b,−6,−7,−13, and−15), chemokines (IL-8, grow regulated alpha protein 1 (GRO)-a, -b, and -g, monocyte chemoattractant protein (MCP)-2 and−4, macrophage inflammatory protein (MIP)-1a and−3a, human beta C-C chemokine-4 (HCC-4), eotaxin, eotaxin-3, thymus-expressed chemokine [TECK, also known as C-C motif chemokine ligand-25 (CCL-25)], C-X-C motif chemokine-5 (CXCL-5 or ENA78), CCL-1 (or I-309), CXCL-11 (or I-TAC), growth and angiogenic factors [such as amphiregulin, angiogenin, epiregulin, heregulin, epidermal growth factor (EGF), basic fibroblast growth factor (bFGF), hepatocyte growth factor (HGF), insulin-like growth factor binding proteins (IGFBP)-2,−3,−4,−6, and−7, keratinocyte growth factor (KGF), nerve growth factor (NGF), placenta growth factor (PIGF), stem cell factor (SCF), stroma cell-derived factor-1 (SDF-1), vascular endothelial growth factor (VEGF)], matrix metalloproteinases (MMP)-1,−3,−10,−12,−13, and−14, metallopeptidase inhibitor (TIMP)-1 and−2, plasminogen activator inhibitor (PAI)-1 and−2, tissue plasminogen activator (tPA), urokinase-type plasminogen activator (uPA); and cathepsin B, receptors/ligands [EGF receptor, Fas ligand, intercellular adhesion molecule (ICAM)-1 and−3, osteoprotegerin (OPG), uPA receptor, soluble gp130 protein (SGP130), soluble tumor necrosis factor receptors (sTNFRs including sTNFR-I and sTNFR-II, and decoy receptor 1 (DCR-1, also known as TRAIL-R3), non-protein molecules (including nitric oxide (NO), prostaglandin E2 (PGE2); and reactive oxygen species (ROS)], and insoluble factors (collagens, fibronectin, and laminin), all of which constitute SASP (22, 23) (Figure 1). Senescent cells undergoing SASP have high metabolic activity (2427). Although the consequence of SASP can be multifarious, the induction of SASP does not depend on the type of triggers such as ROS, DNA damage, oncologic signaling, or cell types (24). First report of SASP was described in human fibroblasts undergoing RS, which showed a strong inflammatory response by using microarray analysis (22). SASP components including IL-6 and−12, MIP-2, and interferon-gamma (IFN-g) were similar between RS and SIPS fibroblasts, suggesting that SASP can be induced by both RS and SIPS (Figure 1) (22, 24). By inducing SASP, senescent cells communicate with immune cells playing a role in their own death, through recruitment of T cells, macrophages, and natural killer cells, which function collectively to clear the senescent cells. To maintain tissue homeostasis, the removal of senescent cells in a timely manner is crucial. With aging, the immune response declines, a phenomenon known as “immunosenescence” (28). As a result, the clearance of senescent cells is impaired.

SASP cells also can recruit myeloid derived suppressor cells (MDSCs), a heterogeneous and immature population of myeloid cells that can suppress immune responses, to prostate and liver tumors and accelerate tumorigenesis. For example, CCL-2, an important SASP factor secreted by senescent cells, attracts MDSCs to the tumor site. In the presence of tumor derived factors, MDSCs fail to differentiate and inhibit the function of other immune cells, such as T cells, dendritic cells, macrophages, and natural killer cells, and thereby creating an immune tolerant environment (2831).

In addition to communicating with immune cells, by secreting extracellular vesicles (EVs), senescent cells undergoing SASP also communicate with surrounding cells to promote senescence in these neighboring cells (3236). EVs or small heterogeneous vesicles are secreted from stressed or activated cells as result of cytoskeletal reorganization (37). EVs are characterized into exosomes (nanometers, <120 nm), micro vesicles (or microparticles, 100–500 nm), and apoptotic bodies released upon fragmentation of apoptotic cells (larger size, 500–5,000 nm) (38). EVs contain proteins, lipids, and nucleic acids (mRNA, DNA, and non-coding RNAs such as microRNAs and long non-coding RNAs). EVs are detected in biological fluids, enriched in specific proteins and lipids, and are produced by all cell types. For instance, through cell-cell interactions and secretion of soluble molecules, mesenchymal stem cells (MSCs) exert their functions on surrounding cells. These functions include anti-inflammation, anti-fibrosis, anti-apoptosis, pro-proliferation, and pro-angiogenesis. Among factors that are secreted by MSCs, PGE2, transforming growth factor-β (TGF-β), IL-6, IL-1 receptor antagonist (IL-1RA), tumor necrosis factor (TNF)-inducible gene 6 protein (TSG6), NO produced by inducible NO synthase (iNOS), or kynurenine produced by indoleamine 2. 3-dioxygenase (IDO) are part of the anti-inflammatory secretome. Other molecules, including HGF, FGF, and VEGF are important components of MSC paracrine activity, which are primarily generated within EVs that have a key role in cell-cell communication (39). With aging, the production of EVs is increased, partly via mechanisms dependent on p53 and its downstream target gene tumor suppression-activated pathway 6 (TSAP6). In endothelial cells, compared to the lower passage (passage 4) and non-senescent cells, the higher passage (passage 21) and activated senescent cells produce an increased number of functional small EVs, which may have a role in vascular physiology and disease (37, 40). Microparticles secreted by senescent endothelial cells increase ROS production and enhance the senescence of neighboring endothelial cells. In addition to that, microparticles also increase the expression of cellular senescent markers p21(WAF-1) and p16(INK4a) in endothelial cells (37, 41).

The induction of SASP is a highly heterogeneous, multi-step, and dynamic process, during which the properties of senescent cells continuously evolve and diversify in a context dependent manner (7). The SASP component is associated with the duration of senescence, the type of senescence stimuli, as well as the cellular origin (42). Long-term persistence of SASP and senescent cells has been shown to promote the development of CVD, cancer, aging-related disease, and aging itself (43).

Senescence-Associated Stemness (SAS), A Unique Phenotype of SASP, is Induced by Sips

Senescence is characterized by cell cycle arrest, and therapy-induced senescence has long been the basis for cancer treatments to inhibit cancer cell growth (25). Recently, this conventional concept has been challenged (4446). Milanovic et al. reported that senescent cancer cells induced by chemotherapy can be reprogrammed to acquire “stemness” (SAS: senescence-associated stemness), which allows them to escape senescence-mediated cell cycle arrest (Figure 1). Importantly, these senescent cells, which escape cell cycle arrest, gain an elevated tumor-initiating capacity possessing enhanced clonogenic growth potential compared to cells that have never undergone senescence (45). SASP is different from senescence-mediated cell cycle arrest (47, 48) and death (49, 50). SASP exerts a range of tumorigenic effects, including promoting angiogenesis, invasion, and metastasis (7) and is now considered one of the key mechanisms in the development of chemoresistance (7, 4951). For instance, previous study has reported that through producing WNT16B, chemotherapy-triggered damage of stroma fibroblast promotes therapy resistance (52). Although the exact molecular mechanisms by which SASP induces cancer therapy resistance remain elusive, it is possible that SASP triggers the formation of cancer stem cells and thereby eluding drug treatment and reproducing tumor (53). Another possibility is that SASP causes cancer treatment resistance through inducing MDSC-driven immune-suppression (28). Therefore, the concepts of SASP and SAS largely overlap (Figure 1), but they describe different biological phenomena. After cancer therapy, subsets of senescent cells produce numerous inflammatory cytokines (54), growth factors (23), ROS (5557), and promote cell growth (24, 58), eventually leading to a more aggressive proliferative phenotype (45, 49, 50) through SAS (59, 60). Although SASP can be induced by both RS and SIPS, it remains unclear whether RS can provoke SAS.

Stress Reprograms Cells to SASP, Locking In and Leading to the Long-Term Effects of SASP

Pro-inflammatory stimuli increase inflammation, but these effects are temporary. The uniqueness and probably most important feature of SASP is its long-term effects to the cells. In fibroblast, Coppé et al. have suggested that a large proportion of SASP produced by senescent fibroblasts is irreversible once established. In these cells, the inactivation of p53 can reverse the growth arrest and resume the cell proliferation, despite the low level of p16(INK4a). These findings suggest that SASP might be permanently locked in an irreversible stage by unknown mechanisms that uncoupling senescence-associated cell cycle arrest from the SASP (24).

Oncogene-induced senescence (OIS), characterized by marked epigenetic changes, can promote tumor progression (23). Via evaluating changes in histone modifications during the senescence of the OIS classic cell model, HRASG12V overexpression in IMR90 cells, or RAS (61), Leon et al. observed an increase of active histone H3K79 di- and tri-methylation (H3K79me2/3) marks at the IL1A locus. This increase corresponds to an increase of the H3K79 methyltransferase disruptor of telomeric silencing 1-like (DOT1L) expression. In OIS cells, DOT1L upregulation is required for H3K79me2/3 and IL1A expression. Knockdown of DOT1L during RAS-induced senescence decreases both DOT1L and H3K79me2/3 occupancy at the IL1A locus but not at other SASP loci, and decreases the expression of IL1A mRNA, the epxresison of IL1A at the cellular membrane, as well as the transcription and secretion of SASP downstream factors. Leon et al. also found that the decrease of SASP was not due to the rescue of senescence-associated cell cycle arrest. Although DOT1L can regulate DDR, and H3K79 methylation promotes 53BP1 binding to sites of DNA double-strand breaks, the authors found no marked changes in 53BP1 or γH2AX foci upon DOT1L knockdown. DOT1L knockdown in BRAF-induced senescent cells or pharmacological inhibition of DOT1L in RAS-induced senescent cells inhibited H3K79 methylation and SASP expression while maintaining the senescence-associated cell cycle arrest. These observations indicated that DOT1L regulates SASP through a DDR-independent mechanism, and that DOT1L expression is required for the SASP but is dispensable for other senescent cell phenotypes. Overall, this study suggested DOT1L as an epigenetic regulator of SASP, whose expression is uncoupled from the senescence-associated cell cycle arrest (62).

Various stresses can induce epigenetic changes and thereby leading to persistent changes in gene expression via inducing chromatin alteration (63). Because these stresses also induce SASP, as in the case of OIS presented above, epigenetic regulation can be one of the potential mechanisms that causes the irreversible stage of SASP (64, 65). Therefore, it is possible that epigenetic changes induced by SASP have some impact on the establishment of long-term effects of SASP. In mid-life flies, acetyl-CoA levels are increased with a corresponding increase in histone acetylation resulting in changes in their transcriptomes (66, 67). However, in mammals, the role of epigenetics in SASP remains unclear. For example, IR-triggered epigenetic changes have been extensively studied, but the results are contradictory (68). Therefore, SASP is potentially induced and maintained by various stresses rather than solely by epigenetic changes, which requires further investigation.

SASP in Tumorigenesis: Foe or Ally?

The role of senescence in cancer is highly controversial. SASP can be induced in both cancer and normal cells by stress, oncogenes, or therapy (therapy induced senescence) (69, 70). However, recent studies of senescence have led investigators to consider this process as a double-edged sword for cancer (70). SASP is associated with changes in p16(INK4a), retinoblastoma (Rb), and p53. Via binding to the E2 factor (E2F), Rb blocks the transcription of several E2F targeted genes that are essential for DNA replication. Cyclin and cyclin dependent kinase CDK4/6 phosphorylate Rb, leading to the release of E2F, ultimately leading to cell cycle progression (71). As one of the most widely known markers of senescence, p16(INK4a) binds CDK4/6 and thereby keeping the cells growth arrested at G1 through inhibiting CDK4/6-triggered Rb phosphorylation (72). p53, another important transcription factors, also plays a vital role in the control of cellular senescence. In the inactive form, p53 is bound to Mdm2, an E3 ubiquitin ligase and destined to ubiquitination and proteasomal degradation (73, 74). In response to stress, p53 is phosphorylated, leading to its release from Mdm2 and activation. Activated p53 increases the transcription of several target genes, including p21Cip1 (CDKN1A), known as another potential marker of senescence. p21Cip1 binds and inhibits cyclin dependent kinase 2 (CDK2), resulting in the activation of Rb and cell cycle arrest (Figure 2). Thus, cellular senescence via the activation of p21Cip1, can lead to cell cycle arrest and inhibition of tumorigenesis (70).

FIGURE 2
www.frontiersin.org

Figure 2. SASP in tumorigenesis.

SASP also attracts immune cells to the tumor site for immune clearance of cancer cells. In liver cancer, OIS hepatocytes secrete CCL-2 that attracts CCL-2+ myeloid cells to the tumor site. CCL2+ myeloid cells are de-differentiated to macrophages, engulf cancer cells, and clear them out. Therefore, SASP can also inhibit tumorigenesis by promoting phagocytosis of pre-malignant cells (75, 76).

Importantly, in cancer progression, the effects of SASP on surrounding cells is context dependent. SASP promotes cancer progression by mediating the de-differentiation and division of the neighboring metastatic cancer cells, or triggering an epithelial-to-mesenchymal transition, one hall mark of cancer. Ritschka et al reported that primary mouse keratinocytes treated with the conditioned media from oncogene-transfected senescent cells induce both senescence and stemness markers such as CD34, Lgr6, Prom1, CD44, Ngfr, and Nestin (77). Lluc Mosteiro et al. showed that c-Myc overexpression enhances both stemness and senescence in the neighboring cells by increasing IL-6 production (78). In addition, senescent cancer cells share similar secretory phenotypes with cancer associated fibroblasts, a functionally heterogeneous population of activated fibroblasts that constitutes a major component of tumor stroma (79).

Two sub populations of cancer associated fibroblasts, inflammatory cancer associated fibroblasts and myofibroblastic cancer associated fibroblasts (80), express smooth muscle actin and soluble factors that promote cancer cell motility and progression. The main tumor-promoting factor secreted by cancer associated fibroblasts is CXCL-12, which is also a component of SASP (81). CXCL-12 promotes cancer cell proliferation as well as the angiogenesis (82). Other common factors secreted by cancer associated fibroblasts and senescent cells include SDF-1, GRO-a and -b, IL-8, MCP-1 and−8, all of which contribute to the promotion of cancer progression (8284). Senescent cancer cells also secret factors that can affect cancer associated fibroblasts in a paracrine manner and attract them to tumor sites. Lastly, via secreting matrix metalloproteinases, senescent cells can also restructure the extracellular matrix that can facilitate cancer growth and reduces contact inhibition (22, 85, 86). Therefore, it is possible that SASP cells can escape from cell cycle arrest and promote tumorigenesis, which is reported as SAS. Although this has become a widely studied topic, how SAS is initiated remains largely unclear.

SASP is also induced by various cancer treatments (87). Conventional cancer treatments include chemotherapy, radiotherapy, chemo-radiotherapy, and surgery. Chemotherapy is a type of standard cancer treatments. In the 1960's, the discovery of anthracyclines (daunorubicin, doxorubicin (DOX), epirubicin, idarubicin, mitoxantrone, and valrubicin) in Italy was a breakthrough in oncology. Despite dramatic changes in cancer treatments in subsequent decades, anthracyclines remain the cornerstone of contemporary chemotherapy for various cancers. In the late 1970's, bleomycin, vinblastine, and cisplatin were used in chemotherapy. From the early nineteenth century, cancer treatments include radical, super-radical and ultra-radical surgery. During 1891–1981, radical mastectomy was used in breast cancer treatments. However, in 1981, the use of radical surgery for cancer treatment was disapproved and was reduced as soon as the combination of systemic adjuvant therapy and local surgery was shown to produce similar results. Systemic adjuvant therapies include radiation and cytostatic drugs are required to treat cancer dissemination and metastasis. In the last 60 years, novel cancer treatments have been drastically developed, including targeted therapies using small molecule inhibitors and monoclonal antibodies (MAbs). Recently, two types of immunotherapies have significantly impacted oncology, including Checkpoint inhibitory MAbs and chimeric antigen-specific receptor (CAR)-transfected T-cells (CAR-T cells). Immunotherapy has been shown to produce durable responses in numerous tumor types. Antigen-specific immune responses can be markedly effective, even in late-stage disease. Additionally, two other types of biological therapies, antitumor vaccines, and oncolytic viruses, have been developed. They are physiological and well-tolerated (88).

Anthracyclines are well-known for both effectiveness and cardiotoxicity (89). The destruction of cardiomyocytes causes cardiac dysfunction. Therefore, cardiomyocytes have been considered the major target of anthracyclines (89). Widely used as a prototypical anticancer drug, DOX treatment causes cardiotoxicity and cardiac dysfunction through the mechanism involves the accumulation of ROS and reactive nitrogen species (RNS) in adult cardiac muscle cells. DOX also targets other cardiac cell types such as cardiac progenitor cells and cardiac fibroblasts. For instance, DOX detrimental effects on endogenous cardiac stem cell (CSC) pool was detected, which induces premature senescence. DOX increases ROS production and DNA damage in resident CSCs with the induction of senescence and apoptosis. DOX altered the myocardium of treated patients who exhibited a higher number of CSCs marked by DNA damage and senescence, particularly by the phosphorylated form of histone H2AX and p16INK4a. Beyond the toxicity on cardiomyocytes and other cardiac cell types, recent studies have suggested that other cell types, including endothelial cells, also play a role in the pathogenesis of anthracycline-induced cardiomyopathy (89, 90). There are two major types of cardiotoxicities caused by anthracyclines: acute and chronic forms. The acute cardiotoxicity occurs after a single dose or a single course of treatment, with symptoms developed within 14 days from the end of treatment. The chronic cardiotoxicity can be further divided into the early onset and the late onset. The early onset chronic cardiotoxicity occurs within a year after treatment, shown as a dilated-hypokinetic cardiomyopathy with progressive evolution toward heart failure. The late onset chronic cardiotoxicity occurs after years or decades from the end of treatment. While the acute cardiotoxicity usually is reversible, the two chronic forms are considered irreversible, with a poor prognosis and a limited to heart failure therapy. Although the specific mechanisms of anthracycline-induced cardiotoxicity remain to be fully elucidated, the involvement of ROS production, DNA damage, cellular senescence and cell death, changes in iron metabolism, and Ca2þ signaling has been suggested. During DNA replication, transcription, or recombination, topoisomerase (Top) 2β uncoils DNA filaments, triggers mitochondrial dysfunction, activates cell death pathways and ROS production. Top2β has been shown to play a critical role in anthracycline-mediated cardiotoxicity (89). Similarly, bleomycin, vinblastine, and cisplatin also induced severe side effects.

Advances in cancer treatments have significantly reduced morbidity and increased survival of cancer patients, but the side effects on non-cancer cells have significantly affected the quality of patient's life. Moreover, cancer cells acquire resistance to therapies and progress to become more aggressive (9193). In addition to the cytotoxic or less direct cytostatic cancer therapies, one of attractive strategies for cancer treatments is to provoke cellular senescence, so-called therapy-induced senescence (94). Therapy-induced senescence creates a cytostatic effect and slows down the growth of cancer cells. However, therapy-induced senescence can also induce SASP, which may cause conflicting effects on tumorigenesis as previously noted. Therefore, a critical role of SASP-modulating therapies in cancer treatments has recently been recognized. Currently, there are two major strategies for SASP-modulating therapies. One strategy is to convert senescent cells from tumor-promoting to an anti-tumorigenic phenotype. For example, Toso et al. reported that, the Pten null mice develop Pten-loss-induced cellular senescence, which is characterized by an immunosuppressive SASP that promotes tumorigenesis (95). Alternatively, inhibition of Jak2/Stat3 signaling reprograms the SASP cytokine networks by restoring senescence surveillance and tumor clearance, which enhances chemotherapy efficacy. We listed other candidates of SASP-modulating therapies in Table 1. Another strategy is using senolytic compounds to selectively eliminate SASP cells through inducing apoptosis known as “senolysis” (118). This strategy was based on the observation that in contrast to non-senescent cells, SASP cells can activate survival pathways and therefore are highly resilient to apoptosis (119). These Senescent Cell Anti-Apoptotic Pathways (SCAPs) such as pathways that regulate Caspase-3 (120), Mcl-1, BLC-2, BCL-XL, and BCL-W etc. (121, 122), might be mediated by senescence-associated mitochondrial dysfunction (6). The upregulation of these anti-apoptotic pathways of SASP cells protect them from apoptotic stimuli including serum withdrawal (121), UV damage (122), oxidative stress (123), extrinsic apoptotic inducers (122, 124), and cytotoxic drugs such as staurosporine or thapsigargin (120, 125). Therefore, a majority of senolytic agents target these anti-apoptotic factors including BCL-2 family of proteins (BCL-2, BCL-XL, and BCL-W, Mcl-1), p53-p21Cip1 axis, hypoxia-inducible factor 1-alpha, heat shock protein 90, several receptor tyrosine kinases, and the PI3K/Akt/mTOR pathway (122, 126). Senolysis induced by glutamate metabolic enzyme GLS1 inhibitor selectively eliminates SASP cells from various organs and tissues and aged mice, ameliorates age-related tissue dysfunction and the symptoms of arteriosclerosis and obese diabetes (127). As SASP cells take weeks to reaccumulate, a “hit-and-run” approach can be used to administer these senolytic drugs. The first senolytic drugs discovered including Dasatinib, Quercetin, Fisetin, and Navitoclax transiently cause apoptosis to SASP cells. Using preclinical models, studies have shown that senolytic drugs can delay or prevent cancer, CVD, as well as complications of radiotherapy. Early pilot trials of senolytic drugs suggest they decrease SASP cells and inflammation in humans. Clinical trials of these drugs as single regimen or in combination are beginning or underway (127, 128). However, hitting a single target in SCAPs may increase off-targeting risk. For example, Navitoclax, the well-known BCL-2 inhibitor that hits a single or few SCAP nodes, causes substantial off-targeting effects on non-senescent cells and thereby making that drug “panolytic” (126, 129131). To reduce off-targeting effects and increase the specificity of senolytic drugs, efforts are made to target more than one SCAP nodes, as evident by the using of combination therapy in clinical trials (NCT028749819, NCT02652052, NCT0463124, NCT02848131, NCT04210986, and NCT03675724). The combination therapy of Dasatinib + Quercetin, and Fisetin showed inhibitory effect on senescent cells in vivo and not on non-senescent cells (132, 133). The senolytic efficacy of the combination therapy Dasatinib + Quercetin was further validated in cultured human senescent cells isolated from the abdominal subcutaneous adipose tissue fragments from diabetic and obese patients after surgery. Treatment of the cells with Dasatinib + Quercetin in culture reduced the senescent cells by 70% within 2 days (134138).

TABLE 1
www.frontiersin.org

Table 1. FDA approved anti-cancer drugs inducing therapy-induced senescence.

SASP Instigates CVD

CVD remain the most common age-related diseases worldwide and the leading cause of death in the aged individuals (139, 140). Studies of human samples and mouse models reveal that senescent cardiovascular cells accumulate at the site of the disease cardio-vasculature and leading to atherosclerosis, heart failure, arterial stiffness, and hypertension (43, 141). The persistent senescence of cardiovascular cells leads to CVD, however cardiovascular cell senescence is also required for the maintenance of cardiovascular homeostasis during embryonic development and wound healing (142). Importantly, with nutritional and growth factor deficiency, cardiovascular cells enter reversible quiescence (143). In the normal aging process, senescent cells accumulate in the cardiovascular system and predispose it to aging-related CVD (43). Following the onset of CVD, the microenvironment of the diseased tissues creates more cellular stress, and a second wave of disease-associated senescent cells is produced enhancing the disease process. Cells of the cardiovascular system including cardiomyocytes, endothelial, vascular smooth muscle, and immune cells (144), significantly contribute to the development and progression of CVD.

Cardiac metabolism has an important role in maintaining the heart's function and cardiovascular homeostasis (145). Cardiac aging is associated with a decreased angiogenic capacity (146), an increased fibrosis (147), metabolic maladaptation (148), cardiomyocyte senescence and dysfunction (149), all of which lead to cardiac remodeling and failure (150). Senescent cardiomyocytes exhibit the hallmarks of DNA damage, mitochondria dysfunction, contractile dysfunction, endoplasmic reticulum stress, hypertrophic growth, and SASP. In the heart, the senescence of cardiomyocytes is also regulated by non-cardiomyocyte cells (endothelial cells, fibroblasts, and immune cells). The senescence of cardiomyocytes also leads to phenotypic and functional changes in those non-cardiomyocyte cells and thereby contributing to cardiac aging and pathological remodeling (43, 151, 152). Nevertheless, the molecular mechanisms by which the senescence of cardiomyocytes is induced and regulated, as well as their interaction with the senescence of non-cardiomyocytes remain to be fully studied. Furthermore, how the local microenvironment of the heart, and how chromatin structure remodeling and DDR activated by cardiomyocytes contribute to the senescence of cardiomyocytes are not well-known. More studies are needed to determine the physiological and pathological functions of senescent cardiomyocyte during cardiac development, regeneration, and pathological remodeling, also to understand whether cardiomyocyte senescence has a role in cardiac aging and the related heart failure with preserved ejection fraction.

Endothelial cells form the inner layer of all blood vessels and communicate with the neighboring cells for tissue regeneration, and control low density lipoprotein (LDL) transcytosis and atherogenesis (153). Endothelial cell dysfunction is associated with the development of atherosclerotic plaques, and is tightly linked to endothelial senescence (154). In the atherosclerotic plaques from human coronary arteries, senescent endothelial cells with high beta-galactosidase activity, a marker of senescent cells, were detected (154). In the initial stages of atherogenesis, an increase of ox-LDL retention was observed in the subendothelial spaces. Senescent endothelial cells undergoing SASP express adhesion molecules including VCAM1 and ICAM1 and secrete various cytokines, leading to the recruitment of circulating monocytes, driving monocyte invasion. Consequently, circulating monocytes invade to the subendothelial spaces, take up the ox-LDL and are converted to foam cell macrophages (155). Both foam cells and SASP endothelial cells secrete a plethora of chemoattractant proteins including IL-1α, TNF -α, and MCP-1 which promote more immune cell recruitment and form atherosclerotic plaques. SASP endothelial cells also cause thrombus formation through the activation of PAI-1, a known marker of senescence (156).

Vascular smooth muscle cells also play a significant role in cardiovascular homeostasis. Vascular smooth muscle cell senescence and pro-inflammatory phenotype has been implicated in the development of CVD, progression of atherosclerosis, and an instigator of ischemic heart disease (157, 158). Compared to normal vascular smooth muscle cells, those isolated from human atherosclerosis exhibited a lower level of proliferation (159) and higher expression level of p16INK4, p21Cip1, hypo phosphorylated Rb, and SA-β-gal activity, suggesting the cells are undergoing SASP (160). Vazquez-Padron et al. reported that vascular smooth muscle cells derived from aged thoracic aortas have higher levels of platelet-derived growth factor receptor-alpha and acquire resistant to apoptosis induced by serum starvation or NO (161). Human vascular smooth muscle cells undergoing SASP had an inactivation of Sirt1, were increased in atherosclerosis (162), and vulnerable atherosclerotic plaque (163).

Patients over the age of 60 who have shorter leukocyte telomere length, a cellular senescence marker, showed 3.18-fold higher in mortality rate from heart failure (164). It is becoming clear that immunosenescence contributes to both innate and adaptive immune systems. Senescent T cells can produce many pro-inflammatory cytokines and chemokines and therefore they have pathogenic potential in CVD such as hypertension, atherosclerosis, myocardial infarction, and heart failure (165). In individuals with human immunodeficiency virus infection treated with a combination of antiretroviral therapy, we found that four components of SASP, including (1) telomere shortening-induced DNA damage and the subsequent induction of p53, p16INK4, and p21Cip1; (2) mitochondrial ROS induction; (3) inflammation; and (4) impairment of efferocytosis, were regulated by p90RSK-mediated ERK5 S496 phosphorylation in myeloid cells. We also found a key role of p90RSK-mediated ERK5 S496 phosphorylation in SASP-mediated atherosclerotic plaque formation (166, 167). Since cancer therapy can induce SASP (22, 54, 168), p90RSK-mediated ERK5 S496 phosphorylation may play a role in SASP induced by cancer therapy.

Cardiovascular Risk Factors = Cancer Risk Factors, Because SASP Is Shared?

The clear epidemiological connection between aging, diabetes mellitus, smoking, cancer, and heart failure raised an obvious question about the pathological link among them (169). Ma et al. showed that shorter telomere length in diabetes mellitus patients is probably due to higher ROS production (170). Cigarette smoke extract increases ROS production and subsequently enhances p16 expression in the progenitor endothelial cells, leading to endothelial dysfunction (56, 171). The crucial role of p16 in smoking induced SASP and subsequent lung injury has also been suggested (172). These data suggest that ROS-mediated SASP induction can be a convergent point in the development of cancer and CVDs. As stated above, ROS is pivotal to initiate telomere DNA damage and the subsequent SASP induction. However, because telomere DNA damage-induced SASP is irreversible once established (17), antioxidant therapy may be no longer effective to attenuate SASP during the progression of cancer and CVD. As such, understanding the regulation of SASP is the key to understand not only the interconnections between cancer and CVD, but also age-related diseases such as diabetes mellitus, Alzheimer's disease, cataract, and chronic obstructive pulmonary disease (Figure 3).

FIGURE 3
www.frontiersin.org

Figure 3. SASP as a hinge between cancer and cardiovascular disease. DM, Diabetes mellitus.

Conclusion

The crucial role of senescence in both cancer and CVD is becoming evident. However, the contribution of senescence to the interconnection between cancer and CVD remains unclear. In this review, we discuss the possible involvement of SASP, which can be instigated by various stresses, including cancer therapy, ROS, and pro-inflammatory cytokines, in the establishment of interconnection between cancer and CVDs. Especially, the new concept of senescence-associated stemness, a unique form of SASP, which may have a significant impact on determining the interplay between cancer and CVD, under highly stressful conditions such as cancer therapy, myocardial infarction, and heart failure. Therefore, although the involvement of senescence in cancer and CVD is a kind of old concept, the perspective of senescence is radically changing.

Author Contributions

PB, NTL, and JA wrote a manuscript. ED edited and made critical suggestions for a manuscript. SK, LR, RA, JC, KS, AD, JH, and SL critically read the manuscript and provided suggestions for improving the manuscript contents. All authors contributed to the article and approved the submitted version.

Funding

The research activity related to this review was partially supported by grants from the National Institutes of Health (NIH) to JA (AI-156921).

Conflict of Interest

SL is an Advisory Board member of AstraZeneca, Beyond Spring Pharmaceuticals, STCube Pharmaceuticals.

The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Antonangeli F, Grimsholm O, Rossi MN, Velotti F. Editorial: cellular stress and inflammation: how the immune system drives tissue homeostasis. Front Immunol. (2021) 12:668876. doi: 10.3389/fimmu.2021.668876

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Meizlish ML, Franklin RA, Zhou X, Medzhitov R. Tissue homeostasis and inflammation. Annu Rev Immunol. (2021) 39:557–81. doi: 10.1146/annurev-immunol-061020-053734

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Nakad R, Schumacher B. DNAD. amage response and immune defense: links and mechanisms. Front Genet. (2016) 7:147. doi: 10.3389/fgene.2016.00147

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Huang RX, Zhou PK. DNA. damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal Transduct Target Ther. (2020) 5:60. doi: 10.1038/s41392-020-0150-x

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Jackson SP, Bartek J. The DNA-damage response in human biology and disease. Nature. (2009) 461:1071–8. doi: 10.1038/nature08467

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Kirkland JL, Tchkonia T. Cellular senescence: a translational perspective. EBioMedicine. (2017) 21:21–8. doi: 10.1016/j.ebiom.2017.04.013

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Kumari R, Jat P. Mechanisms of cellular senescence: cell cycle arrest and senescence associated secretory phenotype. Front Cell Dev Biol. (2021) 9:645593. doi: 10.3389/fcell.2021.645593

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Hayflick L, Moorhead PS. The serial cultivation of human diploid cell strains. Exp Cell Res. (1961) 25:585–621. doi: 10.1016/0014-4827(61)90192-6

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Campisi J. The biology of replicative senescence. Eur J Cancer. (1997) 33:703–9. doi: 10.1016/S0959-8049(96)00058-5

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Liu J, Wang L, Wang Z, Liu JP. Roles of telomere biology in cell senescence, replicative and chronological ageing. Cells. (2019) 8:10054. doi: 10.3390/cells8010054

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Caslini C. Transcriptional regulation of telomeric non-coding RNA: implications on telomere biology, replicative senescence and cancer. RNA Biol. (2010) 7:18–22. doi: 10.4161/rna.7.1.10257

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Davis T, Kipling D. Telomeres and telomerase biology in vertebrates: progress towards a non-human model for replicative senescence and ageing. Biogerontology. (2005) 6:371–85. doi: 10.1007/s10522-005-4901-4

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Dimri GP, Campisi J. Molecular and cell biology of replicative senescence. Cold Spring Harb Symp Quant Biol. (1994) 59:67–73. doi: 10.1101/SQB.1994.059.01.010

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Parrinello S, Samper E, Krtolica A, Goldstein J, Melov S, Campisi J. Oxygen sensitivity severely limits the replicative lifespan of murine fibroblasts. Nat Cell Biol. (2003) 5:741–7. doi: 10.1038/ncb1024

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Naka K, Tachibana A, Ikeda K, Motoyama N. Stress-induced premature senescence in hTERT-expressing ataxia telangiectasia fibroblasts. J Biol Chem. (2004) 279:2030–7. doi: 10.1074/jbc.M309457200

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Bennett G, Papamichos-Chronakis M, Peterson CL. DNA repair choice defines a common pathway for recruitment of chromatin regulators. Nat Commun. (2013) 4:2084. doi: 10.1038/ncomms3084

PubMed Abstract | CrossRef Full Text

17. Fumagalli M, Rossiello F, Clerici M, Barozzi S, Cittaro D, Kaplunov JM, et al. Telomeric DNA damage is irreparable and causes persistent DNA-damage-response activation. Nat Cell Biol. (2012) 14:355–65. doi: 10.1038/ncb2466

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Hewitt G, Jurk D, Marques FD, Correia-Melo C, Hardy T, Gackowska A, et al. Telomeres are favoured targets of a persistent DNA damage response in ageing and stress-induced senescence. Nat Commun. (2012) 3:708. doi: 10.1038/ncomms1708

PubMed Abstract | CrossRef Full Text

19. Fairlie J, Harrington L. Enforced telomere elongation increases the sensitivity of human tumour cells to ionizing radiation. DNA Repair. (2015) 25:54–9. doi: 10.1016/j.dnarep.2014.11.005

PubMed Abstract | CrossRef Full Text | Google Scholar

20. de Magalhaes JP, Chainiaux F, Remacle J, Toussaint O. Stress-induced premature senescence in BJ and hTERT-BJ1 human foreskin fibroblasts. FEBS Lett. (2002) 523:157–62. doi: 10.1016/S0014-5793(02)02973-3

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Birch J, Anderson RK, Correia-Melo C, Jurk D, Hewitt G, Marques FM, et al. DNA damage response at telomeres contributes to lung aging and chronic obstructive pulmonary disease. Am J Physiol Lung Cell Mol Physiol. (2015) 309:L1124–37. doi: 10.1152/ajplung.00293.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Coppe JP, Desprez PY, Krtolica A, Campisi J. The senescence-associated secretory phenotype: the dark side of tumor suppression. Annu Rev Pathol. (2010) 5:99–118. doi: 10.1146/annurev-pathol-121808-102144

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Coppe JP, Kauser K, Campisi J, Beausejour CM. Secretion of vascular endothelial growth factor by primary human fibroblasts at senescence. J Biol Chem. (2006) 281:29568–74. doi: 10.1074/jbc.M603307200

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Coppe JP, Patil CK, Rodier F, Sun Y, Munoz DP, Goldstein J, et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. (2008) 6:2853–68. doi: 10.1371/journal.pbio.0060301

PubMed Abstract | CrossRef Full Text

25. Faget DV, Ren Q, Stewart SA. Unmasking senescence: context-dependent effects of SASP in cancer. Nat Rev Cancer. (2019) 19:439–53. doi: 10.1038/s41568-019-0156-2

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Gorgoulis V, Adams PD, Alimonti A, Bennett DC, Bischof O, Bishop C, et al. Cellular senescence: defining a path forward. Cell. (2019) 179:813–27. doi: 10.1016/j.cell.2019.10.005

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Kuilman T, Peeper DS. Senescence-messaging secretome: SMS-ing cellular stress. Nat Rev Cancer. (2009) 9:81–94. doi: 10.1038/nrc2560

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Salminen A, Kaarniranta K, Kauppinen A. Immunosenescence: the potential role of myeloid-derived suppressor cells (MDSC) in age-related immune deficiency. Cell Mol Life Sci. (2019) 76:1901–18. doi: 10.1007/s00018-019-03048-x

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Di Mitri D, Toso A, Chen JJ, Sarti M, Pinton S, Jost TR, et al. Tumour-infiltrating Gr-1+ myeloid cells antagonize senescence in cancer. Nature. (2014) 515:134–7. doi: 10.1038/nature13638

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Eggert T, Wolter K, Ji J, Ma C, Yevsa T, Klotz S, et al. Distinct functions of senescence-associated immune responses in liver tumor surveillance and tumor progression. Cancer Cell. (2016) 30:533–47. doi: 10.1016/j.ccell.2016.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Gorgun GT, Whitehill G, Anderson JL, Hideshima T, Maguire C, Laubach J, et al. Tumor-promoting immune-suppressive myeloid-derived suppressor cells in the multiple myeloma microenvironment in humans. Blood. (2013) 121:2975–87. doi: 10.1182/blood-2012-08-448548

PubMed Abstract | CrossRef Full Text

32. Takasugi M, Okada R, Takahashi A, Virya Chen D, Watanabe S, Hara E. Small extracellular vesicles secreted from senescent cells promote cancer cell proliferation through EphA2. Nat Commun. (2017) 8:15729. doi: 10.1038/ncomms15728

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Acosta JC, Banito A, Wuestefeld T, Georgilis A, Janich P, Morton JP, et al. A complex secretory program orchestrated by the inflammasome controls paracrine senescence. Nat Cell Biol. (2013) 15:978–90. doi: 10.1038/ncb2784

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Wallis R, Mizen H, Bishop CL. The bright and dark side of extracellular vesicles in the senescence-associated secretory phenotype. Mech Ageing Dev. (2020) 189:111263. doi: 10.1016/j.mad.2020.111263

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Dorronsoro A, Santiago FE, Grassi D, Zhang T, Lai RC, McGowan SJ, et al. Mesenchymal stem cell-derived extracellular vesicles reduce senescence and extend health span in mouse models of aging. Aging Cell. (2021) 20:e13337. doi: 10.1111/acel.13337

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Boulestreau J, Maumus M, Rozier P, Jorgensen C, Noel D. Mesenchymal stem cell derived extracellular vesicles in aging. Front Cell Dev Biol. (2020) 8:107. doi: 10.3389/fcell.2020.00107

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Burger D, Kwart DG, Montezano AC, Read NC, Kennedy CRJ, Thompson CS, et al. Microparticles induce cell cycle arrest through redox-sensitive processes in endothelial cells: implications in vascular senescence. J Am Heart Assoc. (2012) 1:1842. doi: 10.1161/JAHA.112.001842

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Alique M, Ramirez-Carracedo R, Bodega G, Carracedo J, Ramirez R. Senescent microvesicles: a novel advance in molecular mechanisms of atherosclerotic calcification. Int J Mol Sci. (2018) 19:72003. doi: 10.3390/ijms19072003

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Harrell CR, Fellabaum C, Jovicic N, Djonov V, Arsenijevic N, Volarevic V. Molecular mechanisms responsible for therapeutic potential of mesenchymal stem cell-derived secretome. Cells. (2019) 8:50467. doi: 10.3390/cells8050467

PubMed Abstract | CrossRef Full Text

40. Riquelme JA, Takov K, Santiago-Fernandez C, Rossello X, Lavandero S, Yellon DM, et al. Increased production of functional small extracellular vesicles in senescent endothelial cells. J Cell Mol Med. (2020) 24:4871–6. doi: 10.1111/jcmm.15047

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Simoncini S, Chateau AL, Robert S, Todorova D, Yzydorzick C, Lacroix R, et al. Biogenesis of pro-senescent microparticles by endothelial colony forming cells from premature neonates is driven by SIRT1-dependent epigenetic regulation of MKK6. Sci Rep. (2017) 7:8277. doi: 10.1038/s41598-017-08883-1

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Childs BG, Durik M, Baker DJ, van Deursen JM. Cellular senescence in aging and age-related disease: from mechanisms to therapy. Nat Med. (2015) 21:1424–35. doi: 10.1038/nm.4000

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Childs BG Li H, van Deursen JM. Senescent cells: a therapeutic target for cardiovascular disease. J Clin Invest. (2018) 128:1217–28. doi: 10.1172/JCI95146

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Milanovic M, Yu Y, Schmitt CA. The senescence-stemness alliance - a cancer-hijacked regeneration principle. Trends Cell Biol. (2018) 28:1049–61. doi: 10.1016/j.tcb.2018.09.001

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Milanovic M, Fan DNY, Belenki D, Dabritz JHM, Zhao Z, Yu Y, et al. Senescence-associated reprogramming promotes cancer stemness. Nature. (2018) 553:96–100. doi: 10.1038/nature25167

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Dou Z, Berger SL. Senescence elicits stemness: a surprising mechanism for cancer relapse. Cell Metab. (2018) 27:710–1. doi: 10.1016/j.cmet.2018.03.009

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Coppe JP, Rodier F, Patil CK, Freund A, Desprez PY, Campisi J. Tumor suppressor and aging biomarker p16(INK4a) induces cellular senescence without the associated inflammatory secretory phenotype. J Biol Chem. (2011) 286:36396–403. doi: 10.1074/jbc.M111.257071

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Ferrand M, Kirsh O, Griveau A, Vindrieux D, Martin N, Defossez PA, et al. Screening of a kinase library reveals novel pro-senescence kinases and their common NF-kappaB-dependent transcriptional program. Aging. (2015) 7:986–1003. doi: 10.18632/aging.100845

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Saleh T, Tyutynuk-Massey L, Cudjoe EK Jr, Idowu MO, Landry JW, Gewirtz DA. Non-cell autonomous effects of the senescence-associated secretory phenotype in cancer therapy. Front Oncol. (2018) 8:164. doi: 10.3389/fonc.2018.00164

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Saleh T, Tyutyunyk-Massey L, Murray GF, Alotaibi MR, Kawale AS, Elsayed Z, et al. Tumor cell escape from therapy-induced senescence. Biochem Pharmacol. (2019) 162:202–12. doi: 10.1016/j.bcp.2018.12.013

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Munoz DP, Yannone SM, Daemen A, Sun Y, Vakar-Lopez F, Kawahara M, et al. Targetable mechanisms driving immunoevasion of persistent senescent cells link chemotherapy-resistant cancer to aging. JCI Insight. (2019) 5:124716. doi: 10.1172/jci.insight.124716

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Johnson LM, Price DK, Figg WD. Treatment-induced secretion of WNT16B promotes tumor growth and acquired resistance to chemotherapy: implications for potential use of inhibitors in cancer treatment. Cancer Biol Ther. (2013) 14:90–1. doi: 10.4161/cbt.22636

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Cahu J, Bustany S, Sola B. Senescence-associated secretory phenotype favors the emergence of cancer stem-like cells. Cell Death Dis. (2012) 3:e446. doi: 10.1038/cddis.2012.183

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Davalos AR, Coppe JP, Campisi J, Desprez PY. Senescent cells as a source of inflammatory factors for tumor progression. Cancer Metastasis Rev. (2010) 29:273–83. doi: 10.1007/s10555-010-9220-9

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Tsolou A, Passos JF, Nelson G, Arai Y, Zglinicki T. ssDNA fragments induce cell senescence by telomere uncapping. Exp Gerontol. (2008) 43:892–9. doi: 10.1016/j.exger.2008.08.043

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Birch J, Barnes PJ, Passos JF. Mitochondria, telomeres and cell senescence: implications for lung ageing and disease. Pharmacol Ther. (2018) 183:34–49. doi: 10.1016/j.pharmthera.2017.10.005

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Birch J, Passos JF. Targeting the SASP to combat ageing: mitochondria as possible intracellular allies? Bioessays. (2017) 39:235. doi: 10.1002/bies.201600235

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Ohanna M, Giuliano S, Bonet C, Imbert V, Hofman V, Zangari J, et al. Senescent cells develop a PARP-1 and nuclear factor-{kappa}B-associated secretome (PNAS). Genes Dev. (2011) 25:1245–61. doi: 10.1101/gad.625811

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Tang HL, Tang HM, Mak KH, Hu S, Wang SS, Wong KM, et al. Cell survival, DNA damage, and oncogenic transformation after a transient and reversible apoptotic response. Mol Biol Cell. (2012) 23:2240–52. doi: 10.1091/mbc.e11-11-0926

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Gong YN, Crawford JC, Heckmann BL, Green DR. To the edge of cell death and back. FEBS J. (2019) 286:430–40. doi: 10.1111/febs.14714

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Serrano M, Lin AW, McCurrach ME, Beach D, Lowe SW. Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell. (1997) 88:593–602. doi: 10.1016/S0092-8674(00)81902-9

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Leon KE, Buj R, Lesko E, Dahl ES, Chen CW, Tangudu NK, et al. DOT1L modulates the senescence-associated secretory phenotype through epigenetic regulation of IL1A. J Cell Biol. (2021) 220:8101. doi: 10.1083/jcb.202008101

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Chen Z, Xiong ZF, Liu X. Research progress on the interaction between circadian clock and early vascular aging. Exp Gerontol. (2021) 146:111241. doi: 10.1016/j.exger.2021.111241

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Zhu X, Chen Z, Shen W, Huang G, Sedivy JM, Wang H, et al. Inflammation, epigenetics, and metabolism converge to cell senescence and ageing: the regulation and intervention. Signal Transduct Target Ther. (2021) 6:245. doi: 10.1038/s41392-021-00646-9

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Nacarelli T, Liu P, Zhang R. Epigenetic basis of cellular senescence and its implications in aging. Genes. (2017) 8:120343. doi: 10.3390/genes8120343

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Peleg S, Feller C, Ladurner AG, Imhof A. The metabolic impact on histone acetylation and transcription in ageing. Trends Biochem Sci. (2016) 41:700–11. doi: 10.1016/j.tibs.2016.05.008

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Bradshaw PC. Acetyl-CoA metabolism and histone acetylation in the regulation of aging and lifespan. Antioxidants. (2021) 10:572. doi: 10.3390/antiox10040572

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Miousse IR, Kutanzi KR, Koturbash I. Effects of ionizing radiation on DNA methylation: from experimental biology to clinical applications. Int J Radiat Biol. (2017) 93:457–69. doi: 10.1080/09553002.2017.1287454

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Chambers CR, Ritchie S, Pereira BA, Timpson P. Overcoming the senescence-associated secretory phenotype (SASP): a complex mechanism of resistance in the treatment of cancer. Mol Oncol. (2021). doi: 10.1002/1878-0261.13042

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Birch J, Gil J. Senescence and the SASP: many therapeutic avenues. Genes Dev. (2020) 34:1565–76. doi: 10.1101/gad.343129.120

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Topacio BR, Zatulovskiy E, Cristea S, Xie S, Tambo CS, Rubin SM, et al. Cyclin D-Cdk4,6 drives cell-cycle progression via the retinoblastoma protein's C-terminal helix. Mol Cell. (2019) 74:758–70 e4. doi: 10.1016/j.molcel.2019.03.020

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Krishnamurthty J, Torrice C, Ramsey MR, Kovalev GI, Al-Regaiey K, Su LS, et al. Ink4a/Arf expression is a biomarker of aging. J Clin Investig. (2004) 114:1299–307. doi: 10.1172/JCI22475

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Love IM, Shi D, Grossman SR. p53 Ubiquitination and proteasomal degradation. Methods Mol Biol. (2013) 962:63–73. doi: 10.1007/978-1-62703-236-0_5

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Asher G, Shaul Y. p53 proteasomal degradation: poly-ubiquitination is not the whole story. Cell Cycle. (2005) 4:1015–8. doi: 10.4161/cc.4.8.1900

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Huda N, Liu G, Hong H, Yan S, Khambu B, Yin XM. Hepatic senescence, the good and the bad. World J Gastroenterol. (2019) 25:5069–81. doi: 10.3748/wjg.v25.i34.5069

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Liu P, Tang Q, Chen M, Chen W, Lu Y, Liu Z, et al. Hepatocellular senescence: immunosurveillance and future senescence-induced therapy in hepatocellular carcinoma. Front Oncol. (2020) 10:589908. doi: 10.3389/fonc.2020.589908

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Ritschka B, Storer M, Mas A, Heinzmann F, Ortells MC, Morton JP, et al. The senescence-associated secretory phenotype induces cellular plasticity and tissue regeneration. Gene Dev. (2017) 31:172–83. doi: 10.1101/gad.290635.116

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Mosteiro L, Pantoja C, Alcazar N, Marion RM, Chondronasiou D, Rovira M, et al. Tissue damage and senescence provide critical signals for cellular reprogramming in vivo. Science. (2016) 354:aaf4445. doi: 10.1126/science.aaf4445

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Xiang H, Ramil CP, Hai J, Zhang C, Wang H, Watkins AA, et al. Cancer-associated fibroblasts promote immunosuppression by inducing ROS-generating monocytic MDSCs in lung squamous cell carcinoma. Cancer Immunol Res. (2020) 8:436–50. doi: 10.1158/2326-6066.CIR-19-0507

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Ohlund D, Handly-Santana A, Biffi G, Elyada E, Almeida AS, Ponz-Sarvise M, et al. Distinct populations of inflammatory fibroblasts and myofibroblasts in pancreatic cancer. J Exp Med. (2017) 214:579–96. doi: 10.1084/jem.20162024

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Schosserer M, Grillari J, Breitenbach M. The dual role of cellular senescence in developing tumors and their response to cancer therapy. Front Oncol. (2017) 7:278. doi: 10.3389/fonc.2017.00278

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Orimo A, Gupta PB, Sgroi DC, Arenzana-Seisdedos F, Delaunay T, Naeem R, et al. Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell. (2005) 121:335–48. doi: 10.1016/j.cell.2005.02.034

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Begley L, Monteleon C, Shah RB, MacDonald JW, Macoska JA. CXCL12 overexpression and secretion by aging fibroblasts enhance human prostate epithelial proliferation in vitro. Aging Cell. (2005) 4:291–8. doi: 10.1111/j.1474-9726.2005.00173.x

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Wang T, Notta F, Navab R, Joseph J, Ibrahimov E, Xu J, et al. Senescent carcinoma-associated fibroblasts upregulate IL8 to enhance prometastatic phenotypes. Mol Cancer Res. (2017) 15:3–14. doi: 10.1158/1541-7786.MCR-16-0192

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Freitas-Rodriguez S, Folgueras AR, Lopez-Otin C. The role of matrix metalloproteinases in aging: tissue remodeling and beyond. Biochim Biophys Acta Mol Cell Res. (2017) 1864:2015–25. doi: 10.1016/j.bbamcr.2017.05.007

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Quintero-Fabian S, Arreola R, Becerril-Villanueva E, Torres-Romero JC, Arana-Argaez V, Lara-Riegos J, et al. Role of matrix metalloproteinases in angiogenesis and cancer. Front Oncol. (2019) 9:1370. doi: 10.3389/fonc.2019.01370

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Cuollo L, Antonangeli F, Santoni A, Soriani A. The senescence-associated secretory phenotype (SASP) in the challenging future of cancer therapy and age-related diseases. Biology. (2020) 9:120485. doi: 10.3390/biology9120485

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Schirrmacher V. From chemotherapy to biological therapy: a review of novel concepts to reduce the side effects of systemic cancer treatment (Review). Int J Oncol. (2019) 54:407–19. doi: 10.3892/ijo.2018.4661

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Cardinale D, Iacopo F, Cipolla CM. Cardiotoxicity of anthracyclines. Front Cardiovasc Med. (2020) 7:26. doi: 10.3389/fcvm.2020.00026

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Narezkina A, Nasim K. Anthracycline cardiotoxicity. Circ Heart Fail. (2019) 12:e005910. doi: 10.1161/CIRCHEARTFAILURE.119.005910

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Pommier Y, Sordet O, Antony S, Hayward RL, Kohn KW. Apoptosis defects and chemotherapy resistance: molecular interaction maps and networks. Oncogene. (2004) 23:2934–49. doi: 10.1038/sj.onc.1207515

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Longley DB, Johnston PG. Molecular mechanisms of drug resistance. J Pathol. (2005) 205:275–92. doi: 10.1002/path.1706

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Housman G, Byler S, Heerboth S, Lapinska K, Longacre M, Snyder N, et al. Drug resistance in cancer: an overview. Cancers. (2014) 6:1769–92. doi: 10.3390/cancers6031769

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Ewald JA, Desotelle JA, Wilding G, Jarrard DF. Therapy-induced senescence in cancer. J Natl Cancer Inst. (2010) 102:1536–46. doi: 10.1093/jnci/djq364

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Toso A, Revandkar A, Di Mitri D, Guccini I, Proietti M, Sarti M, et al. Enhancing chemotherapy efficacy in Pten-deficient prostate tumors by activating the senescence-associated antitumor immunity. Cell Rep. (2014) 9:75–89. doi: 10.1016/j.celrep.2014.08.044

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Elmore LW, Rehder CW Di X, McChesney PA, Jackson-Cook CK, Gewirtz DA, et al. Adriamycin-induced senescence in breast tumor cells involves functional p53 and telomere dysfunction. J Biol Chem. (2002) 277:35509–15. doi: 10.1074/jbc.M205477200

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Su D, Zhu S, Han X, Feng Y, Huang H, Ren G, et al. BMP4-Smad signaling pathway mediates adriamycin-induced premature senescence in lung cancer cells. J Biol Chem. (2009) 284:12153–64. doi: 10.1074/jbc.M807930200

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Chang BD, Xuan Y, Broude EV, Zhu H, Schott B, Fang J, et al. Role of p53 and p21waf1/cip1 in senescence-like terminal proliferation arrest induced in human tumor cells by chemotherapeutic drugs. Oncogene. (1999) 18:4808–18. doi: 10.1038/sj.onc.1203078

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Mansilla S, Pina B, Portugal J. Daunorubicin-induced variations in gene transcription: commitment to proliferation arrest, senescence and apoptosis. Biochem J. (2003) 372:703–11. doi: 10.1042/bj20021950

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Nagano T, Nakano M, Nakashima A, Onishi K, Yamao S, Enari M, et al. Identification of cellular senescence-specific genes by comparative transcriptomics. Sci Rep. (2016) 6:31758. doi: 10.1038/srep31758

PubMed Abstract | CrossRef Full Text

101. Yosef R, Pilpel N, Papismadov N, Gal H, Ovadya Y, Vadai E, et al. p21 maintains senescent cell viability under persistent DNA damage response by restraining JNK and caspase signaling. EMBO J. (2017) 36:2280–95. doi: 10.15252/embj.201695553

PubMed Abstract | CrossRef Full Text | Google Scholar

102. Yang H, Wang H, Ren J, Chen Q, Chen ZJ. cGAS is essential for cellular senescence. Proc Natl Acad Sci USA. (2017) 114:E4612–E20. doi: 10.1073/pnas.1705499114

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Zhao H, Halicka HD, Traganos F, Jorgensen E, Darzynkiewicz Z. New biomarkers probing depth of cell senescence assessed by laser scanning cytometry. Cytometry A. (2010) 77:999–1007. doi: 10.1002/cyto.a.20983

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Probin V, Wang Y, Zhou D. Busulfan-induced senescence is dependent on ROS production upstream of the MAPK pathway. Free Radic Biol Med. (2007) 42:1858–65. doi: 10.1016/j.freeradbiomed.2007.03.020

PubMed Abstract | CrossRef Full Text | Google Scholar

105. Mei Q, Li F, Quan H, Liu Y, Xu H. Busulfan inhibits growth of human osteosarcoma through miR-200 family microRNAs in vitro and in vivo. Cancer Sci. (2014) 105:755–62. doi: 10.1111/cas.12436

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Probin V, Wang Y, Bai A, Zhou D. Busulfan selectively induces cellular senescence but not apoptosis in WI38 fibroblasts via a p53-independent but extracellular signal-regulated kinase-p38 mitogen-activated protein kinase-dependent mechanism. J Pharmacol Exp Ther. (2006) 319:551–60. doi: 10.1124/jpet.106.107771

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Meng A, Wang Y, Van Zant G, Zhou D. Ionizing radiation and busulfan induce premature senescence in murine bone marrow hematopoietic cells. Cancer Res. (2003) 63:5414–9.

PubMed Abstract | Google Scholar

108. Bu X, Le C, Jia F, Guo X, Zhang L, Zhang B, et al. Synergistic effect of mTOR inhibitor rapamycin and fluorouracil in inducing apoptosis and cell senescence in hepatocarcinoma cells. Cancer Biol Ther. (2008) 7:392–6. doi: 10.4161/cbt.7.3.5366

PubMed Abstract | CrossRef Full Text | Google Scholar

109. Fleury H, Malaquin N, Tu V, Gilbert S, Martinez A, Olivier MA, et al. Exploiting interconnected synthetic lethal interactions between PARP inhibition and cancer cell reversible senescence. Nat Commun. (2019) 10:2556. doi: 10.1038/s41467-019-10460-1

PubMed Abstract | CrossRef Full Text | Google Scholar

110. Alotaibi M, Sharma K, Saleh T, Povirk LF, Hendrickson EA, Gewirtz DA. Radiosensitization by PARP inhibition in DNA repair proficient and deficient tumor cells: proliferative recovery in senescent cells. Radiat Res. (2016) 185:229–45. doi: 10.1667/RR14202.1

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Chatterjee P, Choudhary GS, Sharma A, Singh K, Heston WD, Ciezki J, et al. PARP inhibition sensitizes to low dose-rate radiation TMPRSS2-ERG fusion gene-expressing and PTEN-deficient prostate cancer cells. PLoS ONE. (2013) 8:e60408. doi: 10.1371/journal.pone.0060408

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Morelli MB, Amantini C, Nabissi M, Cardinali C, Santoni M, Bernardini G, et al. Axitinib induces senescence-associated cell death and necrosis in glioma cell lines: The proteasome inhibitor, bortezomib, potentiates axitinib-induced cytotoxicity in a p21(Waf/Cip1) dependent manner. Oncotarget. (2017) 8:3380–95. doi: 10.18632/oncotarget.13769

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Dabritz JH Yu Y, Milanovic M, Schonlein M, Rosenfeldt MT, Dorr JR, et al. CD20-targeting immunotherapy promotes cellular senescence in B-cell lymphoma. Mol Cancer Ther. (2016) 15:1074–81. doi: 10.1158/1535-7163.MCT-15-0627

PubMed Abstract | CrossRef Full Text | Google Scholar

114. Decaup E, Jean C, Laurent C, Gravelle P, Fruchon S, Capilla F, et al. Anti-tumor activity of obinutuzumab and rituximab in a follicular lymphoma 3D model. Blood Cancer J. (2013) 3:e131. doi: 10.1038/bcj.2013.32

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Rosemblit C, Datta J, Lowenfeld L, Xu S, Basu A, Kodumudi K, et al. Oncodriver inhibition and CD4(+) Th1 cytokines cooperate through Stat1 activation to induce tumor senescence and apoptosis in HER2+ and triple negative breast cancer: implications for combining immune and targeted therapies. Oncotarget. (2018) 9:23058–77. doi: 10.18632/oncotarget.25208

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Hasan MR, Ho SH, Owen DA, Tai IT. Inhibition of VEGF induces cellular senescence in colorectal cancer cells. Int J Cancer. (2011) 129:2115–23. doi: 10.1002/ijc.26179

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Schottler J, Randoll N, Lucius R, Caliebe A, Roider J, Klettner A. Long-term treatment with anti-VEGF does not induce cell aging in primary retinal pigment epithelium. Exp Eye Res. (2018) 171:1–11. doi: 10.1016/j.exer.2018.03.002

PubMed Abstract | CrossRef Full Text | Google Scholar

118. Soto-Gamez A, Quax WJ, Demaria M. Regulation of survival networks in senescent cells: from mechanisms to interventions. J Mol Biol. (2019) 431:2629–43. doi: 10.1016/j.jmb.2019.05.036

PubMed Abstract | CrossRef Full Text | Google Scholar

119. Zhu Y, Tchkonia T, Pirtskhalava T, Gower AC, Ding H, Giorgadze N, et al. The Achilles' heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell. (2015) 14:644–58. doi: 10.1111/acel.12344

PubMed Abstract | CrossRef Full Text | Google Scholar

120. Marcotte R, Lacelle C, Wang E. Senescent fibroblasts resist apoptosis by downregulating caspase-3. Mech Ageing Dev. (2004) 125:777–83. doi: 10.1016/j.mad.2004.07.007

PubMed Abstract | CrossRef Full Text | Google Scholar

121. Wang E. Senescent human fibroblasts resist programmed cell death, and failure to suppress bcl2 is involved. Cancer Res. (1995) 55:2284–92.

PubMed Abstract | Google Scholar

122. Yosef R, Pilpel N, Tokarsky-Amiel R, Biran A, Ovadya Y, Cohen S, et al. Directed elimination of senescent cells by inhibition of BCL-W and BCL-XL. Nat Commun. (2016) 7:11190. doi: 10.1038/ncomms11190

PubMed Abstract | CrossRef Full Text

123. Sasaki M, Kumazaki T, Takano H, Nishiyama M, Mitsui Y. Senescent cells are resistant to death despite low Bcl-2 level. Mech Ageing Dev. (2001) 122:1695–706. doi: 10.1016/S0047-6374(01)00281-0

PubMed Abstract | CrossRef Full Text | Google Scholar

124. Sagiv A, Biran A, Yon M, Simon J, Lowe SW, Krizhanovsky V. Granule exocytosis mediates immune surveillance of senescent cells. Oncogene. (2013) 32:1971–7. doi: 10.1038/onc.2012.206

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Tepper CG, Seldin MF, Mudryj M. Fas-mediated apoptosis of proliferating, transiently growth-arrested, and senescent normal human fibroblasts. Exp Cell Res. (2000) 260:9–19. doi: 10.1006/excr.2000.4990

PubMed Abstract | CrossRef Full Text | Google Scholar

126. Chang J, Wang Y, Shao L, Laberge RM, Demaria M, Campisi J, et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat Med. (2016) 22:78–83. doi: 10.1038/nm.4010

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Johmura Y, Yamanaka T, Omori S, Wang TW, Sugiura Y, Matsumoto M, et al. Senolysis by glutaminolysis inhibition ameliorates various age-associated disorders. Science. (2021) 371:265–70. doi: 10.1126/science.abb5916

PubMed Abstract | CrossRef Full Text | Google Scholar

128. Kirkland JL, Tchkonia T. Senolytic drugs: from discovery to translation. J Intern Med. (2020) 288:518–36. doi: 10.1111/joim.13141

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Zhu Y, Tchkonia T, Fuhrmann-Stroissnigg H, Dai HM, Ling YY, Stout MB, et al. Identification of a novel senolytic agent, navitoclax, targeting the Bcl-2 family of anti-apoptotic factors. Aging Cell. (2016) 15:428–35. doi: 10.1111/acel.12445

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Wilson WH, O'Connor OA, Czuczman MS, LaCasce AS, Gerecitano JF, Leonard JP, et al. Navitoclax, a targeted high-affinity inhibitor of BCL-2, in lymphoid malignancies: a phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. (2010) 11:1149–59. doi: 10.1016/S1470-2045(10)70261-8

PubMed Abstract | CrossRef Full Text | Google Scholar

131. Vilgelm AE, Pawlikowski JS, Liu Y, Hawkins OE, Davis TA, Smith J, et al. Mdm2 and aurora kinase a inhibitors synergize to block melanoma growth by driving apoptosis and immune clearance of tumor cells. Cancer Res. (2015) 75:181–93. doi: 10.1158/0008-5472.CAN-14-2405

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Xu M, Pirtskhalava T, Farr JN, Weigand BM, Palmer AK, Weivoda MM, et al. Senolytics improve physical function and increase lifespan in old age. Nat Med. (2018) 24:1246–56. doi: 10.1038/s41591-018-0092-9

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Yousefzadeh MJ, Zhu Y, McGowan SJ, Angelini L, Fuhrmann-Stroissnigg H, Xu M, et al. Fisetin is a senotherapeutic that extends health and lifespan. EBioMedicine. (2018) 36:18–28. doi: 10.1016/j.ebiom.2018.09.015

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Palmer AK, Gustafson B, Kirkland JL, Smith U. Cellular senescence: at the nexus between ageing and diabetes. Diabetologia. (2019) 62:1835–41. doi: 10.1007/s00125-019-4934-x

PubMed Abstract | CrossRef Full Text | Google Scholar

135. Palmer AK, Kirkland JL. Aging and adipose tissue: potential interventions for diabetes and regenerative medicine. Exp Gerontol. (2016) 86:97–105. doi: 10.1016/j.exger.2016.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

136. Palmer AK, Tchkonia T, LeBrasseur NK, Chini EN, Xu M, Kirkland JL. Cellular senescence in type 2 diabetes: a therapeutic opportunity. Diabetes. (2015) 64:2289–98. doi: 10.2337/db14-1820

PubMed Abstract | CrossRef Full Text | Google Scholar

137. Palmer AK, Xu M, Zhu Y, Pirtskhalava T, Weivoda MM, Hachfeld CM, et al. Targeting senescent cells alleviates obesity-induced metabolic dysfunction. Aging Cell. (2019) 18:e12950. doi: 10.1111/acel.12950

PubMed Abstract | CrossRef Full Text | Google Scholar

138. Minamino T, Orimo M, Shimizu I, Kunieda T, Yokoyama M, Ito T, et al. A crucial role for adipose tissue p53 in the regulation of insulin resistance. Nat Med. (2009) 15:1082–7. doi: 10.1038/nm.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

139. Partridge L, Deelen J, Slagboom PE. Facing up to the global challenges of ageing. Nature. (2018) 561:45–56. doi: 10.1038/s41586-018-0457-8

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Roth GA, Johnson C, Abajobir A, Abd-Allah F, Abera SF, Abyu G, et al. Global, Regional, and national burden of cardiovascular diseases for 10 causes, 1990 to 2015. J Am Coll Cardiol. (2017) 70:1–25. doi: 10.1016/j.jacc.2017.04.052

PubMed Abstract | CrossRef Full Text | Google Scholar

141. Song P, Zhao Q, Zou MH. Targeting senescent cells to attenuate cardiovascular disease progression. Ageing Res Rev. (2020) 60:101072. doi: 10.1016/j.arr.2020.101072

PubMed Abstract | CrossRef Full Text | Google Scholar

142. Iop L, Dal Sasso E, Schirone L, Forte M, Peruzzi M, Cavarretta E, et al. The light and shadow of senescence and inflammation in cardiovascular pathology and regenerative medicine. Mediators Inflamm. (2017) 2017:7953486. doi: 10.1155/2017/7953486

PubMed Abstract | CrossRef Full Text | Google Scholar

143. Blagosklonny MV. Cell cycle arrest is not senescence. Aging-Us. (2011) 3:94–101. doi: 10.18632/aging.100281

PubMed Abstract | CrossRef Full Text | Google Scholar

144. Litvinukova M, Talavera-Lopez C, Maatz H, Reichart D, Worth CL, Lindberg EL, et al. Cells of the adult human heart. Nature. (2020) 588:466–72. doi: 10.1038/s41586-020-2797-4

PubMed Abstract | CrossRef Full Text

145. Kolwicz SC Jr, Purohit S, Tian R. Cardiac metabolism and its interactions with contraction, growth, and survival of cardiomyocytes. Circ Res. (2013) 113:603–16. doi: 10.1161/CIRCRESAHA.113.302095

PubMed Abstract | CrossRef Full Text | Google Scholar

146. Moriya J, Minamino T. Angiogenesis, cancer, and vascular aging. Front Cardiovasc Med. (2017) 4:65. doi: 10.3389/fcvm.2017.00065

PubMed Abstract | CrossRef Full Text | Google Scholar

147. Biernacka A, Frangogiannis NG. Aging and cardiac fibrosis. Aging Dis. (2011) 2:158–73.

Google Scholar

148. Costantino S, Paneni F, Cosentino F. Ageing, metabolism and cardiovascular disease. J Physiol. (2016) 594:2061–73. doi: 10.1113/JP270538

PubMed Abstract | CrossRef Full Text | Google Scholar

149. Shimizu I, Minamino T. Cellular senescence in cardiac diseases. J Cardiol. (2019) 74:313–9. doi: 10.1016/j.jjcc.2019.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

150. Lopaschuk GD, Jaswal JS. Energy metabolic phenotype of the cardiomyocyte during development, differentiation, and postnatal maturation. J Cardiovasc Pharmacol. (2010) 56:130–40. doi: 10.1097/FJC.0b013e3181e74a14

PubMed Abstract | CrossRef Full Text | Google Scholar

151. Tang X, Li PH, Chen HZ. Cardiomyocyte senescence and cellular communications within myocardial microenvironments. Front Endocrinol. (2020) 11:280. doi: 10.3389/fendo.2020.00280

PubMed Abstract | CrossRef Full Text | Google Scholar

152. Cardoso AC, Lam NT, Savla JJ, Nakada Y, Pereira AHM, Elnwasany A, et al. Mitochondrial substrate utilization regulates cardiomyocyte cell cycle progression. Nat Metab. (2020) 2:167–78. doi: 10.1038/s42255-020-0169-x

PubMed Abstract | CrossRef Full Text | Google Scholar

153. Huang L, Chambliss KL, Gao X, Yuhanna IS, Behling-Kelly E, Bergaya S. SR-B1 drives endothelial cell LDL transcytosis via DOCK4 to promote atherosclerosis. Nature. (2019) 569:565–9. doi: 10.1038/s41586-019-1140-4

PubMed Abstract | CrossRef Full Text | Google Scholar

154. Minamino T, Miyauchi H, Yoshida T, Ishida Y, Yoshida H, Komuro I. Endothelial cell senescence in human atherosclerosis: role of telomere in endothelial dysfunction. Circulation. (2002) 105:1541–4. doi: 10.1161/01.CIR.0000013836.85741.17

PubMed Abstract | CrossRef Full Text | Google Scholar

155. Weber C, Noels H. Atherosclerosis: current pathogenesis and therapeutic options. Nat Med. (2011) 17:1410–22. doi: 10.1038/nm.2538

PubMed Abstract | CrossRef Full Text | Google Scholar

156. McDonald AP, Meier TR, Hawley AE, Thibert JN, Farris DM, Wrobleski SK, et al. Aging is associated with impaired thrombus resolution in a mouse model of stasis induced thrombosis. Thromb Res. (2010) 125:72–8. doi: 10.1016/j.thromres.2009.06.005

PubMed Abstract | CrossRef Full Text | Google Scholar

157. Zhuge Y, Zhang J, Qian F, Wen Z, Niu C, Xu K, et al. Role of smooth muscle cells in cardiovascular disease. Int J Biol Sci. (2020) 16:2741–51. doi: 10.7150/ijbs.49871

PubMed Abstract | CrossRef Full Text | Google Scholar

158. Stojanovic SD, Fuchs M, Kunz M, Xiao K, Just A, Pich A, et al. Inflammatory drivers of cardiovascular disease: molecular characterization of senescent coronary vascular smooth muscle cells. Front Physiol. (2020) 11:520. doi: 10.3389/fphys.2020.00520

PubMed Abstract | CrossRef Full Text | Google Scholar

159. Bennett MR, Macdonald K, Chan SW, Boyle JJ, Weissberg PL. Cooperative interactions between RB and p53 regulate cell proliferation, cell senescence, and apoptosis in human vascular smooth muscle cells from atherosclerotic plaques. Circ Res. (1998) 82:704–12. doi: 10.1161/01.RES.82.6.704

PubMed Abstract | CrossRef Full Text | Google Scholar

160. Gorenne I, Kavurma M, Scott S, Bennett M. Vascular smooth muscle cell senescence in atherosclerosis. Cardiovasc Res. (2006) 72:9–17. doi: 10.1016/j.cardiores.2006.06.004

PubMed Abstract | CrossRef Full Text | Google Scholar

161. Vazquez-Padron RI, Lasko D, Li S, Louis L, Pestana IA, Pang MH, et al. Aging exacerbates neointimal formation, and increases proliferation and reduces susceptibility to apoptosis of vascular smooth muscle cells in mice. J Vasc Surg. (2004) 40:1199–207. doi: 10.1016/j.jvs.2004.08.034

PubMed Abstract | CrossRef Full Text | Google Scholar

162. Gorenne I, Kumar S, Gray K, Figg N, Yu H, Mercer J, et al. Vascular smooth muscle cell sirtuin 1 protects against DNA damage and inhibits atherosclerosis. Circulation. (2013) 127:386–96. doi: 10.1161/CIRCULATIONAHA.112.124404

PubMed Abstract | CrossRef Full Text | Google Scholar

163. Wang J, Uryga AK, Reinhold J, Figg N, Baker L, Finigan A, et al. Vascular smooth muscle cell senescence promotes atherosclerosis and features of plaque vulnerability. Circulation. (2015) 132:1909–19. doi: 10.1161/CIRCULATIONAHA.115.016457

PubMed Abstract | CrossRef Full Text | Google Scholar

164. Cawthon RM, Smith KR, O'Brien E, Sivatchenko A, Kerber RA. Association between telomere length in blood and mortality in people aged 60 years or older. Lancet. (2003) 361:393–5. doi: 10.1016/S0140-6736(03)12384-7

PubMed Abstract | CrossRef Full Text | Google Scholar

165. Akboga MK, Canpolat U, Yayla C, Ozcan F, Ozeke O, Topaloglu S, et al. Association of platelet to lymphocyte ratio with inflammation and severity of coronary atherosclerosis in patients with stable coronary artery disease. Angiology. (2016) 67:89–95. doi: 10.1177/0003319715583186

PubMed Abstract | CrossRef Full Text | Google Scholar

166. Singh MV, Kotla S, Le NT, Ae Ko K, Heo KS, Wang Y, et al. Senescent phenotype induced by p90RSK-NRF2 signaling sensitizes monocytes and macrophages to oxidative stress in HIV-positive individuals. Circulation. (2019) 139:1199–216. doi: 10.1161/CIRCULATIONAHA.118.036232

PubMed Abstract | CrossRef Full Text | Google Scholar

167. Le NT, Heo KS, Takei Y, Lee H, Woo CH, Chang E, et al. A crucial role for p90RSK-mediated reduction of ERK5 transcriptional activity in endothelial dysfunction and atherosclerosis. Circulation. (2013) 127:486–99. doi: 10.1161/CIRCULATIONAHA.112.116988

PubMed Abstract | CrossRef Full Text | Google Scholar

168. Zhang S, Liu X, Bawa-Khalfe T, Lu LS Lyu YL, Liu LF, et al. Identification of the molecular basis of doxorubicin-induced cardiotoxicity. Nat Med. (2012) 18:1639–42. doi: 10.1038/nm.2919

PubMed Abstract | CrossRef Full Text | Google Scholar

169. Leon BM, Maddox TM. Diabetes and cardiovascular disease: epidemiology, biological mechanisms, treatment recommendations and future research. World J Diabetes. (2015) 6:1246–58. doi: 10.4239/wjd.v6.i13.1246

PubMed Abstract | CrossRef Full Text | Google Scholar

170. Ma D, Zhu W, Hu S, Yu X, Yang Y. Association between oxidative stress and telomere length in Type 1 and Type 2 diabetic patients. J Endocrinol Invest. (2013) 36:1032–7.

PubMed Abstract | Google Scholar

171. He Z, Chen Y, Hou C, He W, Chen P. Cigarette smoke extract changes expression of endothelial nitric oxide synthase (eNOS) and p16(INK4a) and is related to endothelial progenitor cell dysfunction. Med Sci Monit. (2017) 23:3224–31. doi: 10.12659/MSM.902746

PubMed Abstract | CrossRef Full Text | Google Scholar

172. Cottage CT, Peterson N, Kearley J, Berlin A, Xiong X, Huntley A, et al. Targeting p16-induced senescence prevents cigarette smoke-induced emphysema by promoting IGF1/Akt1 signaling in mice. Commun Biol. (2019) 2:307. doi: 10.1038/s42003-019-0532-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: SASP, senescence associated secretory phenotype, cancer, cardiovascular disease, replicative senescence (RS), stress-induced premature senescence (SIPS)

Citation: Banerjee P, Kotla S, Reddy Velatooru L, Abe RJ, Davis EA, Cooke JP, Schadler K, Deswal A, Herrmann J, Lin SH, Abe J and Le NT (2021) Senescence-Associated Secretory Phenotype as a Hinge Between Cardiovascular Diseases and Cancer. Front. Cardiovasc. Med. 8:763930. doi: 10.3389/fcvm.2021.763930

Received: 24 August 2021; Accepted: 16 September 2021;
Published: 20 October 2021.

Edited by:

Alessandra Ghigo, University of Turin, Italy

Reviewed by:

Stefano Ratti, University of Bologna, Italy
Melanie Ricke-Hoch, Hannover Medical School, Germany

Copyright © 2021 Banerjee, Kotla, Reddy Velatooru, Abe, Davis, Cooke, Schadler, Deswal, Herrmann, Lin, Abe and Le. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jun-ichi Abe, jabe@mdanderson.org; Nhat-Tu Le, nhle@houstonmethodist.org

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.