Skip to main content

Cellular senescence or stemness: hypoxia flips the coin

Abstract

Cellular senescence is a complex physiological state whose main feature is proliferative arrest. Cellular senescence can be considered the reverse of cell immortalization and continuous tumor growth. However, cellular senescence has many physiological functions beyond being a putative tumor suppressive trait. It remains unknown whether low levels of oxygen or hypoxia, which is a feature of every tissue in the organism, modulate cellular senescence, altering its capacity to suppress the limitation of proliferation. It has been observed that the lifespan of mammalian primary cells is increased under low oxygen conditions. Additionally, hypoxia promotes self-renewal and pluripotency maintenance in adult and embryonic stem cells (ESCs), induced pluripotent stem cells (iPSCs) and cancer stem cells (CSCs). In this study, we discuss the role of hypoxia facilitating senescence bypass during malignant transformation and acquisition of stemness properties, which all contribute to tumor development and cancer disease aggressiveness.

Background

Cellular senescence is a complex physiological state of the cell, whose main features are proliferative arrest, changes in chromatin structure, specific epigenetic modifications, metabolic changes, high acid β-galactosidase activity, an increased secretion phenotype and organelle alterations. Senescence can be induced by intrinsic and extrinsic stimuli such as telomere shortening and DNA damage response, oncogenic signaling detection, oxidative stress, or anticancer therapies [1, 2]. Cellular senescence is necessary for the regulation of different physiological processes, such as tumor suppression, development of embryo structures, wound repair and cellular reprogramming in tissue repair [3,4,5]. Furthermore, cellular senescence correlates with tissue aging and is involved in the deleterious effects of the diseases associated with it [6].

Despite the physiological relevance of senescence, it is still unknown whether hypoxia, which is a feature of every tissue in the organism, modulates its initiation regulating its full normal development. Thus, it is unclear whether hypoxia works as a stress-inducing physiologically regulated senescence or is able to promote proliferation and immortalization prior to malignant transformation. It has been shown that hypoxia contributes to tumor malignancy [7]. Moreover, hypoxia supports pluripotency maintenance and self-renewal in embryonic stem cells (ESCs), adult stem cells, induced pluripotent stem cells (iPSCs) and cancer stem cells (CSCs) [8,9,10]. In this review, we bring together senescence and hypoxia and study their relationship with immortalization, acquisition of stemness and malignant properties.

Main text

Main features of cellular senescence

Cell cycle arrest

The cell cycle is a process divided into four phases (G1, G2, S and M). Progression from one phase to the next depends mainly on the action of cyclins and cyclin-dependent kinases (CDKs). Ten different cyclins that form four different groups (A, B, D and E) are specifically synthesized and degraded during each phase of the cell cycle, regulating the activity of CDKs. In general, senescent cells present G1-DNA (2N) because of the expression of CDK inhibitors (CKIs) belonging to the INK4 and CIP/KIP families. These CKIs inhibit CDK2, CDK4 and CDK6, which control the progression from G1 to S, and CDK1 in G2/M. Furthermore, during G1 phase of the cycle, there is a checkpoint, known as the restriction point or R point that reunites all of the intrinsic and extrinsic signals to allow or prevent transition to S phase. When cellular senescence is induced, several factors regulate retinoblastoma protein (pRB) to block the G1/S transition. Among these factors, the high expression of CKIs is very important [11, 12].

Classically, cell cycle stop associated with cellular senescence has been considered irreversible. However, some authors support the possibility of return to proliferation under specific genetic or transcriptomic conditions [13]. During replicative senescence, inactivation of p53 allows proliferation if p16INK4a expression is low [14]. The punctual loss of pRB proteins allows cell cycle progression in RAS-induced senescence. It has also been shown that the H3K9 demethylating proteins JMJD2C and LSD1 are able to promote proliferation after oncogene-induced senescence (OIS) [15].

Apoptosis resistance

Senescent cells are also characterized by survival signaling, which promotes resistance to apoptosis. Under normal conditions, severe DNA damage and other kinds of stress can promote the proapoptotic activity of p53 and the expression of its transcriptional targets, NOXA and BAX [16]. However, senescent cells show high activity of antiapoptotic molecules such as BCL-2, BCL-xL, BCL-w, PI3K and p21CIP1 [1, 2]. p21Cip1 is essential for survival signaling, as it has been shown that p21Cip1-deficient mice accumulate DNA damage and induce apoptosis [17]. Additionally, in vivo inhibition of BCL-W and BCL-XL can eliminate senescent cells in the lung and epidermis [18].

Chromatin changes

Accumulated DNA damage in senescent cells has consequences on chromatin structure. Thus, nuclear foci denominating DNA segments with chromatin alterations reinforcing senescence (DNA-SCARS) can be permanently found in senescent cells. Although these DNA-SCARSs contain active DNA repair proteins, such as CHK2 or p53, as well as other transitory DNA repair foci, there are no single strand DNAs, DNAs in synthesis or PML nuclear bodies, and they lack RAP and RAD5 proteins. These foci can promote the senescence-associated secretory phenotype (SASP) and stop the cell cycle. To do this, the foci need p53 and pRB [19]. In senescent cells, it is also possible to observe senescence-associated heterochromatin foci (SAHFs). These structures are located in the promoters of cell cycle genes regulated by E2F, silencing them in collaboration with pRB [20, 21]. SAHFs are generated by reorganization of epigenetic repressor modifications such as H3K9, HP1 or macroH2A without an H1 linker [22, 23]. Another chromatin alteration observed in senescence is satellite distensions associated with senescence (SADSs), which consist of decondensation of the constitutive heterochromatin formed by the pericentromeric satellite DNA. This alteration appears prior to SAHFs and differs from them in that it conserves epigenetic canonical marks and is independent of senescence-specific signaling pathways [24].

Epigenetic modifications

Cells suffer great changes during acquisition of a senescent phenotype. These alterations result in changes in accessibility to chromatin and, accordingly, affect gene expression. DNA methylation and histone modifications are altered during cellular senescence [25].

Replicative senescence is associated with globally hypomethylated DNA in CpG sites, except for specific sites that are hypermethylated. This hypomethylation is associated with deregulation of the DNMT1 enzyme acquired during successive DNA replications and is associated with an increase in p16INK4a and p21CIP1. However, in premature senescence, such as that induced by doxorubicin treatment, radiation or oncogenic RAS overexpression, DNA hypomethylation is not observed [25,26,27,28].

Histones suffer posttranslational modifications in their amino-terminal tail, which changes their interaction with nucleosomes and participates in the regulation of chromatin structure [29]. Replicative senescence is associated with a decrease in global histones and, consequently, reduced modified histones. Thus, it has been suggested that the repressor Rap1 leaves its binding sites at chromosome ends to bind histone coding genes. Additionally, a general decrease in the histone modifications H4K16ac, H3K4me3, H3K9me3 and H3k27me3 and a general increase in H3K9ac and H4K20me3 along with p-H2AX have been reported. Specifically, p-H2AX is associated with telomere shortening and colocalizes with double string break (DBS) repair machinery [25].

Additionally, it has been shown that macroH2A1 increases PARP1 activity, promoting SASP via the PARP/NFkB pathway. In this case, a positive feedback loop is generated because SASP factors promote macroH2A1 expression [30, 31]. At this respect, macroH2A1 presence can be determinant to senescence in tumors as it plays an important role in regulation of SASP and it modulates CSC identity and chemoresistance as it has been described, especially, in hepatocellular carcinoma [32,33,34,35]. However, this role of macroH2A1 is not exclusive to CSCs, as it is implied in stem cell fate and reprogramming [36,37,38,39,40,41] among others. MacroH2A1 is one of the chromatin regulators linked to many of the processes linked such as stemness, senescence, SAPS, hypoxia or DNA damage, being involved in all of them. However, the specificity of the signal regulating the specific mechanism altering each physiological process is still poorly known.

Histone modifications can recruit chromatin remodeling enzymes, which use ATP hydrolysis to change nucleosome positions [29]. This process has also been associated with cellular senescence. For example, ARID1B, which is part of the remodeling complex SWI/SNF, can produce DNA damage and induce reactive oxygen species (ROS) production. In this way, p53, p21CIP1 and p16INK4a expression is increased, and senescence is initiated [42, 43]. Furthermore, the SWI/SNF complex facilitates Polycomb repressive complex 2 (PCR2) evacuation, which can activate the expression of genes encoded by the INK4a/ARF locus [44].

Metabolic changes

Despite their nonproliferative state, senescent cells present high metabolic activity. This is due to a high demand for energy and cellular components to satisfy the senescent phenotype [1, 45, 46]. Catabolism in senescent cells focuses on high consumption of glucose to generate ATP. Glycolysis increases with oncogenic or genotoxic stimuli and during replicative senescence. Furthermore, the presence of dysfunctional mitochondria, which is a common feature of cellular senescence, affects the production of ATP and the NAD+/NADH ratio, which drives proliferative arrest. Additionally, AMPK and NFkB signaling are commonly altered catabolic pathways in senescent cells [45, 47, 48].

On the other hand, during cell senescence, anabolism is focused on the synthesis of proteins, contributing to SASP, and lipids to form new membranous organelles and stimulate autophagy. Glycogen synthesis is also increased in human senescent fibroblasts and during the aging process in different tissues. Additionally, replicative stress is notable in the DNA synthesis process. Related to these anabolic processes, the signaling pathways most altered during cellular senescence are GSK3, ATM, SREBP1, and mTOR [45, 46, 49].

Senescence associated secretory phenotype (SASP)

The senescent-messaging secretome is formed by the pool of cytokines, chemokines and proteases synthesized and secreted by senescent cells. These molecules help cells communicate with surrounding tissues. SASP is a heterogenic and pleiotropic phenotype that depends on genetic context, cell type and microenvironment. Therefore, its functions are also diverse and include positive and negative physiological aspects [1, 50].

The main functions of SASP are autocrine and paracrine reinforcement of the cell cycle stop and an increase in the senescent phenotype. In this process, molecules such as the interleukins IL6 and IL8 or receptors such as IL6R and CXCR2 form a positive feedback loop to promote cellular senescence. This signaling also supposes a tumor suppressor mechanism that stops the growth of nontransformed cells [51,52,53]. SASP is relevant in tissue remodeling, senescence during embryo development, tissue repair and cellular plasticity [5]. Furthermore, senescent cells can recruit the immune system, contributing to the elimination of premalignant and senescent cells. This is possible because of the secretion of proinflammatory interleukins [54].

SASP also contributes to tumor development through the secretion of factors such as VEGF, which promotes angiogenesis or recruits immature myeloid cells. This produces an immunosuppressant effect that is relevant in the development of some tumors, such as prostate or liver [55,56,57]. On the other hand, the proinflammatory activity of the SASP is related to deleterious effects of aging due to the accumulation of senescent cells in tissues. This concept is called “inflammaging” [58].

The functions, composition and regulation of the SASP are also diverse and context-dependent. Nevertheless, there are some key factors that are relevant in this phenotype, such as NFkB, C/EBPB, mTOR, MLLT1, GATA4, p38 MAPK and some effectors of the DNA damage response (DDR). NFkB and C/EBPB constitute a positive feedback loop with IL6 and IL8, which reinforces the secretory phenotype activating NFkB and C/EBPB again [51, 53]. Multiple pathways regulating SASP converge in mTOR, which can promote the action of IL1a, a key factor in the development of SASP. Furthermore, the diversity of secreted molecules can be classified into two secretomes: the inflammatory secretome, which is mainly regulated by IL1, and the TGFB secretome, which depends on yuxtacrine signaling of NOTCH. Although most of the molecules participating in SASP are soluble proteins secreted into the extracellular medium, there are also transmembrane proteins that are cleaved and factors secreted in exosome-like vesicles [1, 50, 58].

Structural and morphological alterations in cellular senescence

Changes in autophagy and lysosomes

Autophagy is a cellular process that enables the recycling of cytoplasmic components, including those in vacuoles, and their degradation in lysosomes. Senescent cells present high autophagic activity due to the high energetic demand and abundance of aberrant cellular components. Furthermore, it is thought that autophagy activation can initiate cellular senescence [1, 20]. As an example, inactivation of mTORC1 during nutrient starvation initiates autophagy and the recycling of cytoplasmic material [59]. This is because mTOR can form a compartment called the TOR autophagy spatial coupling compartment (TASCC), which appears during RAS-induced senescence. Under these conditions, mTOR inhibition decreases SASP proteins, IL6 and IL8 [60].

Even though cells need the activation of autophagy for the proper development of SASP, the synthesis and degradation of proteins are unbalanced in senescent cells, as is the case for other cellular components. This, along with other kinds of stress that are frequent in senescence (oxidative stress, accumulated mutations, lack of chaperons), drives the generation of toxic aggregates made of covalently bound lipids and proteins, increasing the senescent phenotype [1, 45].

To maintain homeostasis, autophagy might be accompanied by high lysosomal activity. Therefore, lysosomal content and activation of lysosomal proteins are increased in senescent cells. It is under discussion whether this is due to biogenesis or the accumulation of these organelles. Inside the lysosomes of senescent cells, there is high activity of the enzyme β-galactosidase. This activity is extensively used as a senescence biomarker and is measured using an artificial reactant known as X-Gal (5-bromo-4-chloro-3-indol-β-D-galactopyranoside) dissolved at an acid pH. The β-galactosidase enzyme is translated from one of the transcripts of the GLB1 gene. Although there are some mechanisms proposed for the transcriptional regulation of this gene and its posttranslational modifications, why it is overexpressed in senescent cells has not been determined [2, 20].

Some types of cells, such as tissue mature macrophages or osteoclasts, present high acid β-galactosidase activity, which is the most commonly used senescent marker along with p16INK4a overexpression [61, 62]. Furthermore, there is solid evidence that supports acid β-galactosidase activity as a physiological phenomenon. Skin aging has been associated with the accumulation of senescent cells in this tissue, which shows high acid β-galactosidase activity [63]. Acid β-galactosidase activity has also been observed in OIS in premalignant lesions in in vivo models and patient samples [3]. Additionally, acid β-galactosidase activity is observed in other physiological processes, such as tissue repair during embryonic development [64, 65].

Dysfunctional mitochondria

Dysfunctional mitochondria accumulate in senescent cells, mainly because of aberrant mitophagic activity. Mitochondria in senescent cells suffer morphological and functional changes, and they change fusion-fission dynamics in favor of the fusion process. Additionally, mitochondria in senescent cells change their membrane potential, which is responsible for proton leakage and an increase in ROS, a factor that is crucial in cellular senescence [1, 2, 66].

Senescence-associated morphology

Senescent cells present a flat and enlarged morphology in vitro. This morphology can be a consequence of the overexpression of CAV1 under stress conditions. This gene encodes CAVEOLIN1, a component of the cholesterol-rich microdomains of the plasma membrane. In human epithelial cells under oxidative stress, p38 MAPK promotes transcriptional activation of CAV1, producing morphological changes in the cells [67]. At the morphological level, it is also possible to observe multinucleated senescent cells. These cells have a defective nuclear lamina because LaminB1 is lost when p53 or pRB is activated [68]. Furthermore, it has been demonstrated that LaminB1 can be used as a biomarker for the detection of damaged cells and during skin regeneration in humans [69].

Mechanisms that trigger cellular senescence

Replicative senescence and DNA damage

Replicative senescence was first described decades ago, when Dr. Leonard Hayflick observed that the number of duplications of human diploid fibroblasts was limited [70]. Later, replicative senescence was observed in multiple cell types, including endothelial cells, keratinocytes, lymphocytes, adrenocortical cells, chondrocytes and smooth muscle vascular cells [71].

Replicative senescence is caused by telomere shortening. Telomeres are structures formed by repetitions of the nucleotide sequence ‘TTAGGG’ and the proteins associated with them at the end of the chromosomes [72]. These structures protect chromosomes from damage and fusion with adjacent chromosomes [73]. Furthermore, telomeres shorten in every cellular duplication due to the inability of DNA polymerases to replicate the end of DNA strings [74]. This was demonstrated in 1990 in human fibroblasts, and later, it was detected in vivo in lymphocytes and cells from liver, skin, blood, and colon tissues [71, 75]. This inability to replicate chromosome ends is solved by the action of the enzyme telomerase (TERT). TERT is an inverse transcriptase that can add ‘TTAGGG’ repetitions to the end of chromosomes using an RNA molecule as a template [76]. The ectopic expression of TERT bypasses senescence in human fibroblasts [77]. However, in Tert-deficient mouse embryonic fibroblasts (MEFs), replicative senescence is not induced earlier. After several generations, the offspring of these Tert-deficient mice present critical short telomeres, which produces genomic instability and a premature aging phenotype, including short lifespan, difficulties in responding to stress and an increase in the appearance of malignant lesions [78, 79]. However, reactivation of Tert mitigates this phenotype [80].

Cells expressing telomerase enzymes differ among mammalian species. In humans, TERT is expressed in embryonic stem cells (ESCs), some adult stem cells and a few somatic cells. In mice, by contrast, it is possible to find Tert activity in several types of adult somatic cells, including fibroblasts. Despite this and the fact that telomeres of mice are 5-10 times longer, their lifespan is 30 times shorter [81,82,83]. Thus, in comparative studies in mammals, the length of telomeres negatively correlates with body size and, in contrast, positively correlates with telomerase activity. This could suggest an evolutionary tendency to short telomeres and low telomerase activity in favor of tumoral suppression. In humans, telomere shortening has been associated with stress- and aging-related pathologies, but not in a causal way. For example, some genetic disorders, such as congenic dyskeratosis, produce dysfunctional telomeres and aging-associated symptoms, such as premature lung fibrosis and cirrhosis [71, 84].

Cells can detect telomere shortening and start the DNA damage response (DDR), which leads to the cell cycle stopping to repair the damage. If this damage is irreparable, signaling drives cells into replicative senescence [73]. The DDR can be activated in cells when other DNA perturbations appear, and they are sensed to severe DNA damage, such as double string breaks (DSBs). Some of these perturbations are common under conditions of oxidative stress, oncogenic stress, ionizing radiation and chemotherapy. Then, the cellular response starts with ATM and ATR kinases phosphorylating H2AX histones around DNA damage and recruiting the DNA repairing complex. Another targets of ATM kinase are CHK1, CHK2, the adaptor protein 53BP1 and the regulator of checkpoint MDC1 [2, 6, 85].

Effectors of proliferative arrest on cellular senescence

p53. Different stimuli can lead to the phosphorylation of different residues of the p53 protein and its stabilization. This induces the cell cycle to stop through activation of CKI p21CIP1 and interaction with the p16INK4a/pRB pathway. p21CIP1 can block the activity of CDK2, CDK4 and CDK6, resulting in hypophosphorylated levels of pRB and proliferative arrest [1, 86]. Moreover, p53 protein is regulated by ARF. ARF helps to maintain p53 levels by binding to HDM2 and blocking its interaction with p53 and its degradation [87]. Additionally, p53 can initiate DNA repair and promote apoptosis, cellular senescence and metabolic changes under a wide range of endogenous and exogenous stimuli [88]. p53 activity is essential for replicative senescence in human and mouse fibroblasts. This is not the case for p21CIP1, which is necessary for cellular senescence in humans but not in mice [89]. However, p21CIP1-deficient mice are more predisposed to develop tumors [90]. Moreover, the relevance of p53 as a tumor suppressor is confirmed by its frequency of mutations in most types of human tumors [88].

CDK inhibitors

As mentioned before, CKIs belong to two different families. On the one hand, the INK4 family is composed of p16INK4a (CDKN2A), p15INK4b (CDKN2B), p18INK4c (CDKN2C) and p19INK4d (CDKN2D), which bind specifically to CDK4 and CDK6, producing a halosteric change in them and blocking the binding of type D cyclins. On the other hand, the CIP/KIP family is composed of p21CIP1 (CDKN1A), p27KIP1 (CDKN1B) and p57KIP2 (CDKN1C), which can bind to CDK-cyclin complexes and inhibit them. All of these proteins are important tumor suppressors because of their role in proliferative arrest [91, 92].

Among this group of CKIs, p16INK4a is especially important because of its relationship with pRB. The p16INK4a/pRB pathway responds directly to several stimuli and indirectly to DNA damage and p53 signaling, which can stop the cell cycle [21]. Specifically, pRB works as a tumor suppressor because it joins and inhibits E2F factors, which regulate the transcription of genes necessary for DNA replication and the cell cycle [93].

The levels of p16INK4a expression increase considerably with the number of replications of cells in vitro; its deletion is usual in immortalized cells, and the loss of its function, by punctual mutation or deletion, is one of the most common mutations in human tumors [20, 94]. Furthermore, p16INK4a silencing through methylation of its promoter has been observed in multiple types of human tumors [95].

p16INK4a, p15INK4b and ARF share the same locus, called INK4a/ARF. Although p15INK4b is similar, functionally and structurally, to p16INK4a, the first has its own open reading frame independent of the other two genes. Even though p16INK4a and ARF share 2 of their 3 exons, they do not present any homology in their amino acid sequences because these proteins are coded in alternative reading frames. Their transcription starts in different locations, and they have different first exons. ARF is also a relevant tumor suppressor gene, as it helps to maintain p53 protein levels [92].

Thus, regulation of the INK4a/ARF locus is crucial for cell cycle stopping and cellular senescence. In normal proliferative cells, the INK4a/ARF locus is silenced due to the polycomb group of proteins (PcG) [96]. Alteration of PcG or the loss of function of some of its components, such as CBX7, BMI1 or EZH2, leads to a loss of silencing epigenetic modifications in the locus and, consequently, to the activation of p16INK4a and ARF expression and cellular senescence. Other epigenetic modifiers that act over the locus INK4a/ARF are CTCF, p300, MLL1 and ZRF1 [1, 20, 96]. The INK4a/ARF locus is regulated by other alternative mechanisms, among which transcription factors are also relevant. Some transcriptional activators are Sp1, Ets, AP1 and PPARy, and some repressor factors are TWIST, YB1 or Id1. Additionally, some of these factors serve as guides for epigenetic modifiers [96].

Oncogene-induced senescence

Cellular senescence can be initiated when mitogenic signals are excessive. This supposes an antitumoral mechanism that is activated by the overexpression of some protooncogenes and the loss of function of tumor suppressor genes. The first evidence of OIS was found after ectopic overexpression of an oncogenic version of HRAS (HRASG12v) in normal nonimmortalized cells. These cells stopped proliferating and acquired a phenotype similar to cellular senescence with high expression of p16INK4a and ARF/p53 [97]. RAS is a member of the protein kinase MAPK signaling pathway, which can activate OIS involving p53 and p16INK4a in the response [98]. Furthermore, hyperactive mutated versions of other proteins of this signaling pathway, such as RAF, MEK and BRAF, and many protooncogenes, such as PI3K, AKT, MYC, ERBB2, and p38 MAPK, can induce a mitogenic signal that activates OIS. Similarly, loss of function of tumor suppressor genes, such as PTEN or NFI, initiates senescence [4, 20].

The expression of oncogenes and hyperactivation of mitogenic signals induce DNA damage and start the DDR. In cells overexpressing mutated HRAS, a hyperreplicative phase that ends in the cell cycle stops induced in OIS with partially replicated DNA is observed [99]. This replicative stress induced by oncogenic activation contributes to the genomic instability of tumoral cells. The mechanisms proposed for this are diverse (replication factors or nucleotides decrease, increase or decrease origin firing, resulting in a reduction in the ratio of elongation forks) [100,101,102]. Recently, a study identified and compared the replication origin sites generated before and after the activation of the oncogenes CCNE1 and MYC. Replication origins were located inside of highly expressed genes after overexpression of both oncogenes. Replication forks collapsed in these cases due to the conflict generated between transcription and replication, and it was also associated with DBs and chromosomal rearrangement breakpoints in the cellular model mentioned and in patient samples [102].

Evidence suggesting that OIS is a physiological phenomenon is abundant in the scientific literature and it has been obtained from animal models and tumor samples from patients. For example, acquired and congenic melanocytic nevi (benign lesions derived from melanocytes) show acid β-galactosidase activity, high expression of p16INK4a and absence of MIB1 without proliferative activity [103, 104]. Furthermore, in transgenic mice that overexpress only BRafV600E, proliferative arrest was observed in melanocytes, while if the tumor suppressor Pten was silenced in these mice, melanomas were generated [105, 106]. Similarly, in the lung, a model of inducible expression of oncogene KRas-V12 was used to study neoplastic lesions. Acid β-galactosidase activity, p16INKa expression and SAHFs were observed in adenomas. However, in the few adenocarcinomas generated, these senescence markers were absent [107]. Similarly, a Pten knockout model showed senescence markers in prostate early-stage neoplasia but not in malignant tumors generated after silencing Trp53 [108]. When NRAS was overexpressed in lymphoid cells of mice, OIS evidence was found. Lymphomas were produced when an additional p53 mutation or Suc39h1 (histone methyltransferase enzyme that mediates pRb silencing) mutation was induced along with the oncogenic overexpression of NRAS [109]. Additionally, in the pituitary gland of mice, senescence markers have been found after E2F overexpression, with proliferation and tumorigenesis stopping [110]. Furthermore, microadenomas have been found in this gland in humans with a prevalence of 10-5% with evidence that OIS could stop the development of these lesions [111].

Therapy-induced senescence

Senescence-induced therapy (TIS) is initiated in tumoral cells and other cells of patients as a consequence of radiotherapy, chemotherapy, or targeted therapy against cancer disease [112, 113]. Evidence of TIS has accumulated since the discovery of OIS. Several studies have shown the relevance of p53 and p16INK4a in the induction of senescence in tumors after treatment and its relationship with prognosis in vivo [114,115,116]. This finding broke the paradigm of the immortality of tumoral cells and the belief that they do not senesce [117].

Most treatments against cancer disease produce DNA damage in tumoral and nontumoral cells of patients. This initiates DDR and, possibly, senescence. It has been described that ionizing radiation can induce TIS in a dose- and cell type-dependent manner. At the molecular level, the cell cycle stop induced by ionizing radiation depends on p21CIP1 and the mutational state of p53, as has been shown in breast cancer, colorectal cancer, and glioblastoma [118, 119]. Furthermore, ionizing radiation can increase ROS production in cells, especially mitochondrial ROS, damaging DNA. It is also known that AKT protein can contribute to ROS increase and senescence initiation via p53 after radiotherapy treatment [120]. In vivo, ionizing radiation damages hematopoietic progenitors in bone marrow due to an increase in ROS [121].

Inducing senescence in tumors has positive effects, including proliferative stop and apoptosis in tumoral cells and, therefore, tumor suppression. It has also been suggested that paracrine induction of senescence could be beneficial for tumor treatment [113]. This has been verified in tumoral cells, but their response to SASP-conditioned medium differs among cell lines [122]. However, TIS can induce protumoral effects in the microenvironment. For example, SASP production, specifically IL6 and IL8, can induce a malignant and invasive phenotype in epithelial premalignant cells. Additionally, tumoral cells can exit the senescence phenotype, resulting in relapse in patients. These cells are called dormant cells and have been detected, among others, in prostate, breast, colon, lung cancer and glioblastoma [123,124,125,126,127,128].

Oxidative stress-induced senescence

There are many different theories that connect oxidative stress, cellular senescence and aging. Most of them have been developed since Dr. Denham Harman proposed the free radical theory of aging in 1956. According to this theory, “aging must be related to the attack of free radicals, which are generally products of metabolic processes, to cellular components” [129]. Later, this theory was focused on the production of free radicals in the mitochondria, and it was proposed that free radicals’ deleterious effects on the cell were caused by the prooxidant-antioxidant imbalance, in favor of oxidation [130].

The most common free radicals produced in the normal metabolism of the cell are superoxide anion (O2-) and hydroxyl radical (-OH). Specifically, dysfunctional mitochondria are characterized by electron leakage, which generates O2- as a secondary byproduct, especially in the reactions produced in complex I (NADH dehydrogenase) and complex III (cytochrome bc1 complex). Different kinds of molecules, such as lipids, proteins and DNA, are damaged by these radicals. When oxidative stress reaches high levels over long periods of time, cells acquire a senescence-like phenotype. Furthermore, ROS accumulation seems to contribute to OIS induction, TIS and p16INK4a-mediated senescence [131].

DNA can be damaged by ROS in multiple ways: single-base damage, double string crosslinking, DNA-protein adducts, and double string bounds [132]. Additionally, ROS have been related to telomere shortening [133]. Thus, ROS can initiate DDR and, consequently, cell senescence. It has also been proposed that the existence of a positive feedback loop between DDR and ROS is promoted by the long-term activation of p21CIP1 by p53. Moreover, p53 can activate other prooxidant genes, such as TP53I3 (PIG3); as a consequence, mitochondria are damaged, and electron leakage increases [134, 135].

Proteins are another target of ROS contributing to cellular oxidative stress. ROS play an important role in cellular signaling through oxidation-reduction of amino acids, generally methionine and cysteine residues. However, some oxidative modifications are not enzymatically reversible, such as carbonylation. When these irreversible modifications accumulate in the proteins of the cells and there is an excess of oxidated proteins, cells try to degrade them via proteosomes. Nevertheless, oxidated proteins bound by crosslinking inhibit proteosomes and accumulate in lysosomes [131, 136].

There are several mechanisms in the cell to avoid damage produced by ROS. This regulation of homeostasis is performed by endogenous antioxidant molecules such as glutathione and enzymes such as superoxide dismutases, catalases and thiol peroxidases. Additionally, cells can use exogenous antioxidant molecules such as micronutrients and vitamins [131]. Some studies indicate that a lack of antioxidant enzymes contributes to premature aging. For example, PrxIII-deficient mice show reduced physiological capacities compared to wild-type mice [137]. Reduced expression of PrxI accelerates the appearance of hemolytic anemia and tumors in mice [138]. In human fibroblasts, the expression of sulfoxide reductases is reduced during replicative senescence [139]. Moreover, nonenzymatic antioxidant molecules seem to contribute to senescence delay. The balance of oxidated/reduced glutathione grows in favor of oxidation with age in blood and in stem cells from adipose tissue in humans [131, 140]. However, noncytotoxic high doses of synthetic antioxidant compounds induce premature senescence in mesenchymal stem cells [141].

The role of p53 in responding to oxidative stress depends on ROS levels. On the one hand, p53 promotes the generation of ROS when oxidative stress is persistent. On the other hand, under transient or mild oxidative stress, p53 can activate the transcription of genes with antioxidant functions, such as GPX1, SOD2, TIGAR, SESN1, SESN2 and SLC2A9 [135].

Despite all of the evidence and classic theories, the number of studies that question the deleterious effect of ROS in senescence and aging has recently grown. It has been observed in yeast and C. elegans that an increase in ROS and oxidative stress do not necessarily accelerate aging [142]. In mice, genetic modification to increase mitochondrial ROS production does not accelerate aging, and the increase in antioxidants does not extend the lifespan of the animals [143].

Several compounds and enzymes with antioxidant activity can extend the lifespan of human fibroblasts by reducing ROS and mitochondrial damage and decreasing telomere shortening. Nevertheless, the addition of some of these compounds to the culture medium can increase ROS production. This is explained by the ability of antioxidants to be oxidized. Something similar is reported in studies that show how culture of many types of cells in hypoxia can increase lifespan but also increase mitochondrial ROS production [144]. However, the role of ROS in cellular signaling and homeostasis has gained attention, as it can be implicated in proliferation and cell survival in certain circumstances. For this reason, it has been proposed that even though ROS participate in the signaling of damage caused by aging, perhaps its accumulation with age is not a cause [145].

To understand the dual role of ROS in cells, the mitohormesis theory has been developed as a stress and signaling mediator. This theory suggests that different kinds of cellular stress can initiate mitochondria-related signaling by which cells increase their tolerance to ROS, unfold proteins and other deleterious elements for cell homeostasis. In this response, the activation of antioxidant mechanisms is important [146] (Fig. 1).

Fig. 1
figure 1

General mechanisms of cellular senescence. Cellular senesce can be induced by several stimuli like DNA damage, telomere shortening, oxidative stress, gain of oncogenes and loss of function of tumor suppressor genes, therapy against cancer disease and, additionally, SASP paracrine signaling. Response to these stimuli is mediated by the main effectors of cellular senescence which are p53 and genes translated from INK4a/ARF locus. Consequently, senescent cells develop a complex phenotype, which most representative feature is cell cycle stop. Additionally, other features of cellular senescence are represented

Hypoxia, cellular senescence and immortalization

Mechanisms for physiological adaptation to hypoxia

What is the effect of environmental oxygen? Atmospheric oxygen levels are approximately 20%; however, availability in our organism is not the same. Oxygen should be distributed from the respiratory system to every tissue through blood flow, and consequently, not every cell is exposed to the same oxygen concentration. For example, the oxygen concentration in the human brain is near 4%, similar to skeletal muscles or the liver. The heart presents a value of 3.3%, and the lungs, despite being the organs responsible for oxygen intake, are supplied with oxygen levels of 5.6%. Other organs, such as the smooth intestine or kidneys, show levels higher than 8%, and in the skin, by contrast, oxygen only reaches 1% [147, 148] (Fig. 2).

Fig. 2
figure 2

Oxygen levels in different organs of human body. Oxygen levels diminish after caption from atmosphere by respiratory organs and distribution along circulatory system to different organs. Oxygen concentrations in human organs are expressed in percentages here

There is both a systemic and a molecular response to hypoxia in mammals. The complete physiological response consists of an acute reflex of hyperventilation and sympathetic activation to incorporate more oxygen and distribute throughout the whole organism [149]. To develop this, there are specialized organs containing oxygen-sensitive cells, such as carotid bodies [150, 151]. Any circumstance or pathology that could mean an interruption in the caption and distribution of oxygen can trigger this response. For example, oxygen concentration decreases with geographic altitude, which triggers a systemic adaptive response [152].

Additionally, at the pathological level, failures in breathing, oxygen flux or hemoglobin levels can trigger a response to hypoxia. Some of the diseases that can activate the physiological mechanisms mentioned are chronic obstructive pulmonary disease (COPD), sleep apnea, lung fibrosis, ischemic cardiopathy and chronic cardiac failure. Moreover, it has been demonstrated that other conditions and pathologies, such as hypertension, diabetes, and some hepatic and renal diseases, are also related to hypoxia [149, 153].

As mentioned, there are adaptive mechanisms to hypoxia at the molecular level. The most relevant is the mechanism carried out by hypoxia inducible factors (HIFs). This molecular pathway can be activated in every cell and depends on the stabilization of HIFα subunits (HIF1α, HIF2α or HIF3α) in hypoxia [154]. When this happens, HIFα subunits dimerize with HIF1β, also known as ARNT, and are translocated to the nucleus. Once there, the heterodimer binds hypoxia response elements (HREs) (G/ACGTG) in the DNA and promotes the expression of the target genes [155, 156].

The most important regulation of HIFα protein is posttranslational, and it is performed mainly by prolyl-hydroxylase enzymes (PHDs), which are sensitive to oxygen levels. There are 4 isoforms of these enzymes, among which PHD2 is the most important, and its optimal activation range is under normoxia and moderate hypoxia. PHDs hydrolyze 1 or 2 proline residues that are well conserved among HIFα proteins, making these proteins susceptible to ubiquitination mediated by von Hippel-Lindau protein (VHL). This last posttranslational modification marks HIFα for degradation in proteosomes. Moreover, the factor inhibiting HIF1α (FIH), another oxygen-sensitive protein, also hydrolyzes HIFα residues. In this case, however, modifications are made in asparagine residues and at lower oxygen levels than PHDs. This inhibition is mostly specific to HIF1α because HIF2α is, in comparison, resistant [156, 157].

There are notable functional differences between HIF1α and HIF2α isoforms. While HIF1a is considered a ubiquitous protein, the activity of HIF2α is more specific to different cell types, such as cardiomyocytes, endothelial cells, hepatocytes or glial cells. Furthermore, in many cases, HIF1α has been identified as responsible for the acute response, while HIF2α has been described as responsible for the chronic response and geographic altitude response. Both proteins share some common transcriptional targets, such as GLUT1 or VEGF, but they specifically regulate others, such as LDHA and PGK1, in the case of HIF1α and EPO and MMP9 and OCT4 in the case of HIF2α. These specificities depend on the cellular context and oxygen concentration [158] (Fig. 3). Additionally, there is a third HIFα subunit, called HIF3α, whose expression in adult tissues is located in the thymus, brain, lungs, heart and kidneys. This isoform presents alternative splicing variants, and in some of them, there is no transactivator domain, as there is in HIF1α and HIF2α. For this reason, HIF3α is thought to be an inhibitor of HIF-dependent transcription. In particular, one of the splicing variants, called IPAS, can bind HIF1α and form a heterodimer that is unable to activate the transcription of HRE-containing genes [157, 159].

Fig. 3
figure 3

HIF signaling and effects of hypoxia on stem cells. In normoxia, HIFα proteins can be hydroxilate by PHDs, this allows ubiquitination mediated by VHL and its degradation in proteasome. However, when oxygen levels are low, HIFα proteins are stable, and they can translocate to nucleus where they dimerize with HIF1β and bind to HRE sequences in the DNA. HIF dimers activate transcription of hypoxia response genes. Through this mechanism and others, hypoxia promotes glycolytic flux and expression of stemness-related genes supporting stem cell properties in adult and embryonic stem cells, iPSCs and cancer stem cells. OSKM: Yamanaka factors Oct3/4, Sox2, Klf4 and cMyc

Relationship between molecular response to hypoxia and cellular senescence

Hypoxia is a feature of locally advanced tumors, and a lack of vascularization deprives cells from oxygen. Furthermore, hypoxia promotes tumor progression and is a marker for poor prognosis in patients [7, 9]. However, it is not yet clear whether hypoxia works as a stressor inducing cellular senescence or can facilitate immortalization prior to malignant transformation, contributing to tumorigenesis. In this respect, several studies describe how different oxygen concentrations, especially low levels of oxygen, can alter the lifespan of cells. In the case of human fibroblasts, low oxygen levels extend lifespan by 20%, while bovine fibroblasts increase lifespan by 80% and 500% in fibroblasts derived from mice [160,161,162,163] (Table 1).

Table 1 Lifespan increase in different cellular models cultured under different oxygen concentrations

The increase in the lifespan of MEFs cultured under hypoxia compared to normoxia has been justified by the difference in accumulated oxidative damage. MEFs proliferated in hypoxia despite conserving the wild-type p19ARF/p53 pathway, which is essential for avoiding immortalization [164], and expressing high levels of p16INK4a. Moreover, MEFs from mice deficient in different proteins involved in DNA repair stopped proliferating with elevated signs of DNA damage when they were cultured in 3% O2. However, they continued proliferating when telomerase activity was suppressed. Therefore, it has been suggested that DNA damage accumulation induces senescence in MEFs under normoxia. When the p19ARF/p53 pathway is altered in these conditions, cells become insensitive to this damage and are immortalized. Additionally, the existence of a mutagenic or adaptive response acquired in hypoxia has been suggested, which allows the extension of lifespan despite p19ARF/p53 and p16INK4a activity. This adaptation only appears at a minimum frequency in cells under normoxia compared to hypoxia.

In contrast, it has been suggested that human cells initiate replicative senescence due to telomere shortening [165], and they have a greater capacity to prevent or repair DNA damage. This may explain why the lifespan increase produced by hypoxia in human cells is lower than that produced by MEFs. This hypothesis tries to explain the different susceptibility to cancer and aging speed between the two species [162]. Nevertheless, this extreme sensitivity to oxygen damage seems to be a particular feature of mouse lab strains, perhaps related to very long telomeres. Therefore, there might be alternative mechanisms altering lifespan in mammals, as oxygen sensitivity does not correlate with lifespan in other species [166]. It has also been proposed that an increase in glycolysis contributes to reducing oxidative damage and can facilitate cellular immortalization. An increase in the expression of PGM can bypass replicative senescence and Ras-induced senescence in MEFs due to the increase in glycolytic flux [167]. Alternatively, this lifespan extension has been causally associated with different cellular signaling pathways, especially the HIFα pathway. HIF1α and HIF2α directly target genes that can regulate the cell cycle, such as p21CIP1 and MYC [168, 169]. It has been suggested that HIF1α can activate the expression of hTERT in the presence of mitochondrial ROS [170]. Moreover, p53 inhibition can occur in a HIFα-dependent and HIFα-independent manner [171], and suppression of the conversion from proliferative arrest to irreversible senescence (a processed named geroconversion) can be independent of HIFα and p53 [172].

Among the mechanisms proposed to promote senescence via HIF1α, the cell cycle stops through activation of p21CIP1 and p27KIP1 [173,174,175] or inhibition of CDC25A [176]. Additionally, hypoxia is related to SASP and, consequently, to paracrine and autocrine induction of senescence. HIF1α can activate the transcription of multiple genes that are relevant for SASP, such as IL8, CXCR2, GROα, IL6 and PAI1. Hypoxia has also been associated with senescence independent of HIF proteins. In anoxia, different pathways converge on the cell cycle stop without HIF interaction [177]. In cells with loss of VHL, senescence is induced by p400 and p27KIP1 [178]. Additionally, observation of senescent cells in mouse melanoma is related to high expression of BCL2 in hypoxia [179].

Hypoxia and stem cell capacity

Hypoxia and embryonic and adult stem cells

Hypoxia has also been associated with stem cells and pluripotency. The first evidence found was related to organismal development because embryos are exposed to a partial state of hypoxia during gestation [180]. The mammalian oviductal and uterine lumen are hypoxic (1.5% O2). This means that the preimplemented embryo is adapted to a low oxygen atmosphere. In humans, it has been shown that low oxygen atmospheres can produce high-quality embryos and improve implementation from developing zygotes [181]. Additionally, placental development is associated with hypoxia, as cytotrophoblasts, which are stem cells forming placenta, proliferate better at low oxygen concentrations in vitro [182, 183]. Moreover, it has been demonstrated that embryonic and somatic stem cells reside in a hypoxic niche. This niche generates a microenvironment in which oxygen levels are even lower than those in tissues where they are located (1-8% O2) [184].

ESCs can be maintained in vitro indefinitely when they are cultured under specific conditions. These specific conditions try to activate the main signaling pathways of pluripotency and self-renewal. In ESCs, the activity of the transcription factors Oct3/4, Sox2 and Nanog, which form the core of the pluripotency circuit, is essential. Oct3/4 can bind DNA to control the expression of multiple genes related to pluripotency. Oct3/4 can form a heterodimer with Sox2 and regulate the expression of genes such as Lefty1, Fgf4, Oct3/4, Sox2 and Nanog. Furthermore, Oct3/4 binds to enhancers of pluripotency genes, and its expression in ESCs is tightly regulated because very high or low levels of this protein can induce differentiation. Nanog can work in coordination with Oct3/4 and Sox2, but its expression levels fluctuate in ESC populations. Among its transcriptional targets, it is possible to find genes related to pluripotency, such as Esrrb1, Rif1, Foxd3 and Rest. In addition to these 3 transcription factors, other proteins play an important role in the maintenance of ESC capacities, such as SMAD1 or STAT3, and some coordinators of its activity, such as Klf4, Esrrb, cMyc and Tfcp2 l1. Generally, the action of all of these proteins leads to the activation of signaling pathways such as LIF/Stat3, Wnt/β-Catenin, FGF/ERK, TGF/SMAD and PKC [185].

Multiple studies have demonstrated that hypoxia promotes self-renewal and pluripotency maintenance in ESCs and other types of stem cell cultures. It is known that self-renewal, pluripotency and the expression of stem cell markers and effectors such as Nanog, Oct3/4 and Sox2 are promoted by hypoxia, either dependent or independent of HIFα. For example, it has been shown that low oxygen concentrations promote neural crest and hemopoietic stem cell survival and prevent ESC differentiation. In vitro, hESCs benefit from low oxygen conditions, reducing differentiation and increasing the expression of pluripotency genes. MYC, OCT4, SOX2 and NANOG are upregulated by HIF2a in these conditions, promoting pluripotency. Additionally, it has been shown that HIF2a can increase glycolytic flux under 5% oxygen, upregulating CTBP1 and CTBP2 to promote self-renewal [181].

Despite all evidence generated, the influence of hypoxia on the differentiation and dedifferentiation capacity of these cells is still controversial. This is probably because of the variability generated by multiple factors, such as the maturation state of the stem cells, hypoxia exposure time or oxygen levels used [8, 186,187,188]. Hypoxia culture can promote the differentiation of mESCs toward endothelial cells, simulating the physiological process of embryonic vascular differentiation [181]. Indeed, the pluripotency markers Oct3/4, Tra1-60, Nanog and Sox2 are downregulated under high levels of ROS. This promotes ESC differentiation into mesodermal and endodermal lineages through MAPK signaling pathways. This effect of ROS in ESCs can be counteracted by the addition of antioxidant compounds to the cell culture [189].

Hypoxia and induced pluripotent stem cells

Mammalian cells can be reprogrammed to induce pluripotent stem cells (iPSCs) through transduction of transcription factors Oct3/4, Sox2, Klf4 and cMyc (OSKM), also known as Yamanaka factors [190, 191]. However, the application of this biological resource in cell therapy is currently limited. This is due to low efficiency in the iPSC generation process and the oncogenic risk entailed in reprogramming protocols. This risk involves the activation of protooncogene cMyc and the use of virus, which can alter the genome when exogenous genetic constructs are randomly inserted in receptor cells [192, 193]. It has been demonstrated that hypoxia increases reprogramming efficiency and reduces the number of transcription factors needed to achieve reprogramming, but it is still unknown how this effect is produced [10]. Several studies have proposed that different molecular effectors are responsible, but discussion has been focused on HIFα proteins. Among the mechanisms described, the metabolic shift from mitochondrial oxidation to glycolysis, which is common to stem cells and iPSCs, is relevant. For this metabolic reprogramming, it seems that HIF1α and HIF2α are both necessary and that their activation increases reprogramming efficiency and pluripotency. Nevertheless, the activation of each HIFα factor occurs differently in every reprogramming stage, and stabilization of HIF2α in the final reprogramming stages inhibits iPSC generation [8, 194,195,196].

First, cellular senescence acts as a barrier for cellular reprogramming. This was determined due to the activation of the main effector pathways of cellular senescence (p53, p16INK4a/pRB, p21CIP1, DDR and INK4a locus remodeling) after overexpression of Yamanaka factors. Another important piece of evidence is the silencing of the INK4a/ARF locus in mature iPSCs and ESCs. It has been demonstrated that inhibition of these cellular senescence effectors can increase reprogramming efficiency in human and mouse fibroblasts, although it entails oncogenic risk [197, 198].

Later, it was observed that senescence could promote reprogramming efficiency in vivo via paracrine action of SASP. When Yamanaka factors are overexpressed, the senescence barrier is activated in tissues, and only some cells allow reprogramming. Senescent cells initiate the secretion of SASP factors, including IL6. The secretion of this interleukin and the activation of its downstream targets increase the reprogramming efficiency of nonsenescent surrounding cells. In this signaling process, activation of the INK4a locus, which induces a senescent phenotype with expression of IL6, is necessary. However, it has been demonstrated that p53, Arf and p21CIP1 are not necessary in this process. This increase in reprogramming efficiency induced by senescence is observed in senescence induced by damage to tissues and in aging [199,200,201].

Recently, more evidence about this relationship between senescence and reprogramming has been uncovered. It has been demonstrated that cyclic and continuous overexpression of Yamanaka factors interferes in epigenetic remodeling associated with aging in mouse models [202]. Chemotherapy-induced senescence mediated by p53 and the epigenetic marker H2K9me3 contributes to the acquisition of stem cell properties in tumoral cells in vivo [203]. Furthermore, it has been observed that SASP promotes cell plasticity and tissue regeneration. When senescent cells were inoculated into the livers of mice, liver stem cell markers were increased [204].

Hypoxia and cancer stem cells

Tumors are heterogeneous entities composed of cells with different identities, including cancer stem cells (CSCs). CSCs are characterized by self-renewal and differentiation in other types of tumoral cells. They can initiate a tumor individually. Although CSC and stem cell capacities are similar, the mechanisms regulating these capacities are deregulated in CSCs [205]. CD34+/CD38- cells were first identified in 1997 as unique cells able to regenerate myeloid leukemia in mice [206]. Since then, CSCs have been isolated from breast, colon, brain, and many other tumors [7].

CSCs are not a static identity in tumoral heterogeneity, as they are considered a cellular state. Cells in the tumor are differentiated and dedifferentiated to this cellular state, and the tumor microenvironment plays an essential role. For example, it has been demonstrated that CSCs can influence nearby fibroblasts and transform them into cancer-associated fibroblasts (CAFs). These CAFs can promote the WNT and NOTCH pathways, contributing to CSC identity maintenance [207, 208]. Stromal mesenchymal stem cells promote CSC maintenance through activation of the NFκB signaling pathway and secretion of CXCL12, IL6 and IL8. Additionally, CSCs can modulate the microenvironment secreting TGFβ to mimic the stem cell niche microenvironment, promote epithelial-mesenchymal transition (EMT) and modify distal niches secreting VGFA, TGFβ, TNFα and LOX.

Hypoxia is part of the tumoral microenvironment, and it is relevant in locally advanced tumors in which vascularization is unable to provide proper flow of oxygen and nutrients for every cell. This lack of oxygen initiates a survival response. The hypoxia response contributes to tumor progression, angiogenesis, metabolic reprogramming, modulation of immune response, metastasis, and therapy resistance. For these reasons, hypoxia clinically correlates with tumor aggressiveness and poor disease prognosis. Additionally, mediators of the hypoxia response at the molecular level have been considered therapeutic targets [7, 9].

As mentioned, different signaling pathways participate in the response to this microenvironment condition, but HIFα signaling is the most studied. It is also known that tumors that conserve normal oxygen conditions have some oncogenes, such as AKT, and the loss of some tumor suppressor genes, such as PTEN, PML, and TSC, can also activate translation of HIFα proteins [7]. It has been demonstrated that hypoxia can promote dedifferentiation and CSC property maintenance. Among these genes promoted by hypoxia, OCT3/4, SOX2, NANOG and KLF4 are important, but the NOTCH pathway and epigenetic modifiers such as BMI1 and SIRT1 also contribute to this effect [9, 209,210,211]. Furthermore, it seems that HIF1α can also counteract the cMYC effect, stopping the cell cycle and maintaining CSC identity [173].

There is a direct correlation between dormant CSCs and hypoxia. In the hypoxic and necrotic regions of tumors, cells show a dedifferentiated phenotype and dormant state, while they are present in vascularized regions [125]. Thus, it has been proposed that hypoxia can activate a dormant state to increase cell survival in CSCs exposed to stress conditions. For example, in glioblastoma cells, hypoxia activates PP2A, which mediates the dormant state program, inducing cell cycle arrest in G1/S phase [128]. In the prostate, HIF1α can activate the expression of CXCR4 and NDRG1, which is regulated by nMYC to induce a dormant state [126, 127]. Furthermore, HIGD1A, a transcriptional target of HIF1a, can promote a dormant state through the production of ROS [124]. This dormant state shares some features with cellular senescence like the cell cycle stop and, also a secretome similar to SASP, which includes factors like IL1-a, IL1-b, IL-6, IL-8, CXCL1 and CXCL2. This secretome may contribute to cancer progression [212]. For instance, it has been shown that prostate cancer cells can induce a transient senescence phenotype with proliferative arrest through SPARC secretion. This senescent stay can be considered dormant state and correlates with bad prognosis [213]. There is evidence that CSC secretes different SASPs programs [212, 213]. This could mean that CSCs, which do not experience senescence and are resistant to chemotherapy, could produce a different secretome and escape destruction by the immune system, that remove instead senescent cells from the tissues. A hypoxic microenvironment could, in the same way, influence the secretome, therefore altering the behavior of the immune system in the site.

Furthermore, in physiologically hypoxic environments, senescent cells express lower levels of pro-inflammatory SASP factors [214]. This is due to AMPK activation induced by hypoxia, which in turn lead to AMPK-dependent suppression of the mTOR-NFκB signaling, thus regulating SAPS secretion [214]. Since the mTOR maintained high activity seems important for “geroconversion”, its decreased activity induced by low oxygen could be important in the bypass of senescence by hypoxia. These results could indicate that one pathway, rather than pleiotropic signaling, can control senescence physiology through SAPS. These results underline the important role of SAPS in malignancy and resistance to therapy.

Hypoxia signaling and ROS have been related because both can activate CSC stress signaling through the TGFβ and TNF pathways to promote cell survival and maintenance of CSC identity and, additionally, promote EMT. Furthermore, TGFβ can stabilize HIFα proteins. Moreover, HIF1α can increase the synthesis of glutathione in response to chemotherapy stress, which promotes the acquisition of the CSC phenotype [7].

It has been proposed that hypoxia can alter the CSC population in tumors in two different ways: promoting dedifferentiation of tumor cells and limiting the differentiation of CSCs. In breast primary carcinoma, hypoxia increases the population of CD44+/CD24- and increases ALDH+ through the HIF1α and AKT/β-catenin signaling pathways [215]. However, in ER-negative breast cancer, it seems that PHD3 activates the NFκB pathway, increasing the CD44+/CD24- population in hypoxia [216].

Both isoforms, HIF1α and HIF2α, have important functions in hypoxia and stem cell property acquisition in tumors. Nevertheless, it seems that tumors preferentially express the HIF2α isoform. In glioma cell lines, there is a subpopulation of cells with greater migration capacity. These cells express high levels of SOX2 and OCT3/4, which are induced by HIF2α [217]. Additionally, in gliomas, the cell membrane marker CD44 cleaves and releases its cytoplasmic fraction. This fragment of protein binds HIF2α to increase the expression of stemness-associated genes [218]. In a mouse model of MYC-induced leukemia, NANOG and SOX2 facilitate MYC binding to the HIF2α promoter, which maintains the CSC state through inhibition of p53 and ROS production [219]. Furthermore, HIF2α can activate the expression of LIF, and the expression of both genes is positively correlated in patients with colorectal cancer [220].

It has been demonstrated that EMT provides high aggressiveness to tumoral growth because cells acquire CSC properties, motility, invasive capacity and dissemination, senescence resistance, therapy resistance and resistance to the immune system [221]. Hypoxia favors metastasis, increasing the expression of genes related to EMT, whose expression is frequent in CSCs. Hypoxia can activate the transcription of SNAI1, ZEB1, TWIST1 and TCF3 through HIF1α [222, 223]. In breast cancer, hypoxia promotes the expression of ZEB1, and it inhibits MYB to promote EMT. Furthermore, the expression of EMT genes is necessary for the appearance of circulating tumor cells (CTCs). These cells can cause metastasis and present CSC properties. CTCs with CSC features have been found in the bloodstream of patients with different types of tumors. For example, hepatocarcinoma CTCs are CD45-/ICAM1+, CTCs in prostate cancer are CD133+ and those in breast cancer are ALDH+. In many cases, CTCs found in patients express OCT3/4, SOX2 and NANOG [7]. Moreover, hypoxia can modulate the secretion of vesicles to increase the expression of GTPase associated with endosomes, RAB22A, and the exosome marker CD63. These exosomes in hypoxia usually contain molecules such as VEGFR2, TNFα1, β-catenin, AKT and EGFR and contribute to tumor progression, angiogenesis, immune suppression, invasion, and metastasis [224,225,226].

Tumors show high glycolytic activity in substitution for oxidative phosphorylation even under normal oxygen conditions and despite its lower efficiency. This phenomenon is known as the “Warburg effect” and involves the widespread consumption of glucose as a consequence [227]. This offers an advantage under hypoxia, which is reinforced by the adaptive response mediated by HIF inducing the expression of key glycolytic enzymes and inhibiting mitochondrial metabolism. HIF1α can promote the activation of PDK1, inhibiting the conversion of pyruvate into acetyl-CoA and inhibiting flux to the tricarboxylic acid cycle (TCA). HIF1α can activate the transcription of LDHA, which converts pyruvate into lactate and limits flux to acetyl-CoA. Furthermore, HIF1α promotes the transcription of PKM2, which is the last step in glycolysis, and it also works as a coactivator of HIF1α to promote glycolysis and tumor growth [9]. Transformed cells express alternative isoforms of key glycolytic enzymes compared to normal cells because of HIF1α activation. For example, HIF1α induces the expression of HK2 and ENO1 and helps to inhibit apoptosis and promote cell migration. Moreover, HIF1α can induce the expression of the GLUT glucose transporter [228, 229]. These transporters have been found to be overexpressed in CSCs in glioblastoma, ovarian cancer and pancreas, and they can promote self-renewal and tumor initiation capacities [230].

Hypoxia has also been related to epigenetic modifications that promote tumor progression. For example, it has been shown that hypoxia can modulate TET activity, which induces de novo methylation in the sequence of genes associated with EMT, such as INSIG1 [168, 231]. Hypoxia can modulate the action of histone modifiers such as HDAC3 and WAD5, which mediate EMT in head and neck cancer and in breast tumors [232]. Additionally, miRNAs such as miR-21, miR-200 and miR-210 are expressed in hypoxia and regulate EMT [233,234,235,236]. It is also known that HIFα proteins interfere with the regulation of the expression of ZNF127 and ALKBH5, which can demethylate the 3’-UTR of KLF4 and NANOG mRNA, increasing their expression in breast tumors [237] (Fig 3).

Conclusions

Oxygen levels are heterogeneous in mammalian tissues, and external and internal conditions can produce additional variations. For this reason, mammals have developed physiological and molecular responses to low levels of oxygen. Current knowledge about cellular senescence has been experimentally developed at atmospheric oxygen concentrations. This may be a bias because the cell lifespan can be extended under hypoxic conditions. Although oxygen damage sensitivity has been proposed to be responsible for different lifespan extensions between humans and mice, it is still controversial. There might be alternative mechanisms controlling lifespan extension induced by hypoxia. Similarly, it is still under discussion whether hypoxia promotes differentiation or helps to maintain stem cell identity under certain conditions. Nevertheless, it has been widely demonstrated that culture of ESCs and some adult stem cells under hypoxic conditions helps to maintain self-renewal and pluripotency. Additionally, hypoxia contributes to the CSC pool and correlates with cancer malignancy and poor prognosis. Therefore, we speculate that evidence generally supports the role of hypoxia as a proliferation and dedifferentiation stimulus whose mechanisms should be studied in depth to understand cellular senescence, stem cell biology and cancer disease.

Availability of data and materials

No datasets were generated during the current study.

Abbreviations

CAFs:

Cancer associated fibroblasts

CDKs:

Cyclin dependent kinases

CKIs:

CDK inhibitors

COPD:

Chronic obstructive pulmonary disease

CSCs:

Cancer stem cells

CTCs:

Circulating tumor cells

DDR:

DNA damage response

DNA-SCARS:

DNA segments with chromatin alterations reinforcing senescence

DSBs:

Double string breaks

EMT:

Epithelial-mesenchymal transition

ESCs:

Embryonic stem cells

FIH:

Factor inhibiting HIF1α

HIFs:

Hypoxia inducible factors

HREs:

Hypoxia response elements

iPSCs:

Induced pluripotent stem cells

OIS:

Oncogene induced senescence

OSKM:

YAMANA factors Oct3/4, Sox2, Klf4 and cMyc

PcG:

Polycomb group of proteins

PHDs:

Prolyl-hydroxylases

pRB:

Retinoblastoma protein

PRC2:

Polycomb repressive complex 2

ROS:

Reactive oxygen species

SADSs:

Satellite distensions associated to senescence

SAHFs:

Senescence associated heterochromatin foci

SASP:

Senescence associated secretory phenotype

TASCC:

TOR autophagy spatial coupling compartment

TCA:

Tricarboxylic acid cycle

TIS:

Therapy induced senescence

VHL:

Von Hippel-Lindau protein

References

  1. Herranz N, Gil J. Mechanisms and functions of cellular senescence. J Clin Invest. 2018;128(4):1238–46.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Hernandez-Segura A, Nehme J, Demaria M. Hallmarks of Cellular Senescence. Trends Cell Biol [Internet]. 2018;28(6):436–53 Available from: https://doi.org/10.1016/j.tcb.2018.02.001.

    Article  CAS  PubMed  Google Scholar 

  3. Chandeck C, Mooi WJ. Oncogene-induced cellular senescence. Adv Anat Pathol. 2010;17(1):42–8.

    Article  CAS  PubMed  Google Scholar 

  4. Collado M, Serrano M. Senescence in tumours: Evidence from mice and humans. Nat Rev Cancer [Internet]. 2010;10(1):51–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Rhinn M, Ritschka B, Keyes WM. Cellular senescence in development, regeneration and disease. Dev. 2019;146(20):dev151837.

    Article  CAS  Google Scholar 

  6. van Deursen JM. Senolytic therapies for healthy longevity. Science. 2019;364(6441):636–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Carnero A, Lleonart M. The hypoxic microenvironment: A determinant of cancer stem cell evolution. Insid Cell. 2016;1(2):96–105.

    Article  Google Scholar 

  8. Huang X, Trinh T, Aljoufi A, Broxmeyer HE. Hypoxia Signaling Pathway in Stem Cell Regulation: Good and Evil. Curr Stem Cell Reports. 2018;4(2):149–57.

    Article  CAS  Google Scholar 

  9. Qiu GZ, Jin MZ, Dai JX, Sun W, Feng JH, Jin WL. Reprogramming of the Tumor in the Hypoxic Niche: The Emerging Concept and Associated Therapeutic Strategies. Trends Pharmacol Sci [Internet]. 2017;38(8):669–86.

    Article  CAS  PubMed  Google Scholar 

  10. Yoshida Y, Takahashi K, Okita K, Ichisaka T, Yamanaka S. Hypoxia Enhances the Generation of Induced Pluripotent Stem Cells. Cell Stem Cell. 2009;5(3):237–41.

    Article  CAS  PubMed  Google Scholar 

  11. Malumbres M, Barbacid M. To cycle or not to cycle: A critical decision in cancer. Nat Rev Cancer. 2001;1(3):222–31.

    Article  CAS  PubMed  Google Scholar 

  12. Malumbres M, Barbacid M. Cell cycle, CDKs and cancer: A changing paradigm. Nat Rev Cancer. 2009;9(3):153–66.

    Article  CAS  PubMed  Google Scholar 

  13. Lee S, Schmitt CA. The dynamic nature of senescence in cancer. Nat Cell Biol. 2019;21(1):94–101.

    Article  CAS  PubMed  Google Scholar 

  14. Beauséjour CM, Krtolica A, Galimi F, Narita M, Lowe SW, Yaswen P, et al. Reversal of human cellular senescence: Roles of the p53 and p16 pathways. EMBO J. 2003;22(16):4212–22.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Yu Y, Schleich K, Yue B, Ji S, Lohneis P, Kemper K, et al. Targeting the Senescence-Overriding Cooperative Activity of Structurally Unrelated H3K9 Demethylases in Melanoma. Cancer Cell. 2018;33(2):322–336.e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Moll UM, Zaika A. Nuclear and mitochondrial apoptotic pathways of p53. FEBS Lett. 2001;493(2-3):65–9.

    Article  CAS  PubMed  Google Scholar 

  17. 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(15):2280–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Rodier F, Muñoz DP, Teachenor R, Chu V, Le O, Bhaumik D, et al. DNA-SCARS: Distinct nuclear structures that sustain damage-induced senescence growth arrest and inflammatory cytokine secretion. J Cell Sci. 2011;124(Pt 1):68–81.

    Article  CAS  PubMed  Google Scholar 

  20. Carnero A. Markers of cellular senescence. Methods Mol Biol. 2013;965:63–81.

    Article  CAS  PubMed  Google Scholar 

  21. Campisi J, D’Adda Di Fagagna F. Cellular senescence: When bad things happen to good cells. Nat Rev Mol Cell Biol. 2007;8(9):729–40.

    Article  CAS  PubMed  Google Scholar 

  22. Salama R, Sadaie M, Hoare M, Narita M. Cellular senescence and its effector programs. Genes Dev. 2014;28(2):99–114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Chandra T, Kirschner K, Thuret JY, Pope BD, Ryba T, Newman S, et al. Independence of Repressive Histone Marks and Chromatin Compaction during Senescent Heterochromatic Layer Formation. Mol Cell. 2012;47(2):203–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Swanson EC, Manning B, Zhang H, Lawrence JB. Higher-order unfolding of satellite heterochromatin is a consistent and early event in cell senescence. J Cell Biol. 2013;203(6):929–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Cheng LQ, Zhang ZQ, Chen HZ, Liu DP. Epigenetic regulation in cell senescence. J Mol Med. 2017;95(12):1257–68.

    Article  CAS  PubMed  Google Scholar 

  26. Koch CM, Reck K, Shao K, Lin Q, Joussen S, Ziegler P, et al. Pluripotent stem cells escape from senescenceassociated DNA methylation changes. Genome Res. 2013;23(2):248–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Sakaki M, Ebihara Y, Okamura K, Nakabayashi K, Igarashi A, Matsumoto K, et al. Potential roles of DNA methylation in the initiation and establishment of replicative senescence revealed by array-based methylome and transcriptome analyses. PLoS One. 2017;12(2):e0171431.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Bielak-Zmijewska A, Wnuk M, Przybylska D, Grabowska W, Lewinska A, Alster O, et al. A comparison of replicative senescence and doxorubicin-induced premature senescence of vascular smooth muscle cells isolated from human aorta. Biogerontology. 2014;15(1):47–64.

    Article  CAS  PubMed  Google Scholar 

  29. Bannister AJ, Kouzarides T. Regulation of chromatin by histone modifications. Cell Res. 2011;21(3):381–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Chen H, Ruiz PD, McKimpson WM, Novikov L, Kitsis RN, Gamble MJ. MacroH2A1 and ATM Play Opposing Roles in Paracrine Senescence and the Senescence-Associated Secretory Phenotype. Mol Cell. 2015;59(5):719–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Ohanna M, Giuliano S, Bonet C, Imbert V, Hofman V, Zangari J, et al. Senescent cells develop a parp-1 and nuclear factor-κB-associated secretome (PNAS). Genes Dev. 2011;25(12):1245–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lo Re O, Fusilli C, Rappa F, Van Haele M, Douet J, Pindjakova J, et al. Induction of cancer cell stemness by depletion of macrohistone H2A1 in hepatocellular carcinoma. Hepatol Int. 2018;67(2):636–50.

    Article  CAS  Google Scholar 

  33. Lo Re O, Mazza T, Giallongo S, Sanna P, Rappa F, Luong TV, et al. Loss of histone macroH2A1 in hepatocellular carcinoma cells promotes paracrine-mediated chemoresistance and CD4+CD25+FoxP3+ regulatory T cells activation. Theranostics. 2020;10(2):910–24.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Borghesan M, Fusilli C, Rappa F, Panebianco C, Rizzo G, Oben JA, et al. DNA Hypomethylation and histone variant macroH2A1 synergistically attenuate chemotherapy-induced senescence to promote hepatocellular carcinoma progression. Cancer Res [Internet]. 2016;76(3):594–606 [cited 2021 Jun 21], Available from: https://pubmed.ncbi.nlm.nih.gov/26772755/.

    Article  CAS  Google Scholar 

  35. Lo Re O, Douet J, Buschbeck M, Fusilli C, Pazienza V, Panebianco C, et al. Histone variant macroH2A1 rewires carbohydrate and lipid metabolism of hepatocellular carcinoma cells towards cancer stem cells. Epigenetics [Internet]. 2018;13(8):829–45 [cited 2021 Jun 21], Available from: /pmc/articles/PMC6224214/.

    Article  Google Scholar 

  36. Cedeno RJ, Nakauka-Ddamba A, Yousefi M, Sterling S, Leu NA, Li N, et al. The histone variant macroH2A confers functional robustness to the intestinal stem cell compartment. PLoS One [Internet]. 2017;12(9) [cited 2021 Jun 21], Available from: https://pubmed.ncbi.nlm.nih.gov/28934364/.

  37. Barrero MJ, Sese B, Martí M, JCI B. Macro histone variants are critical for the differentiation of human pluripotent cells. J Biol Chem [Internet]. 2013;288(22):16110–6 [cited 2021 Jun 21], Available from: https://pubmed.ncbi.nlm.nih.gov/23595991/.

    Article  CAS  Google Scholar 

  38. Bereshchenko O, Lo Re O, Nikulenkov F, Flamini S, Kotaskova J, Mazza T, et al. Deficiency and haploinsufficiency of histone macroH2A1.1 in mice recapitulate hematopoietic defects of human myelodysplastic syndrome. Clin Epigenetics [Internet]. 2019;11(1) [cited 2021 Jun 21], Available from: https://pubmed.ncbi.nlm.nih.gov/31439048/.

  39. Gaspar-Maia A, Qadeer ZA, Hasson D, Ratnakumar K, Adrian Leu N, Leroy G, et al. MacroH2A histone variants act as a barrier upon reprogramming towards pluripotency. Nat Commun [Internet]. 2013;4:1565 Available from: https://pubmed.ncbi.nlm.nih.gov/23463008/.

    Article  CAS  Google Scholar 

  40. Giallongo S, Rehakova D, Raffaele M, Lo Re O, Koutna I, Vinciguerra M. Redox and Epigenetics in Human Pluripotent Stem Cells Differentiation [Internet]. Antioxid Redox Signal. Mary Ann Liebert Inc. 2021;34:335–49 [cited 2021 Jun 21]. Available from: https://pubmed.ncbi.nlm.nih.gov/32567336/.

    Article  CAS  Google Scholar 

  41. Pasque V, Radzisheuskaya A, Gillich A, Halley-Stott RP, Panamarova M, Zernicka-Goetz M, et al. Histone variant macroH2A marks embryonic differentiation in vivo and acts as an epigenetic barrier to induced pluripotency. J Cell Sci [Internet]. 2012;125(24):6094–104 [cited 2021 Jun 21]. Available from: https://pubmed.ncbi.nlm.nih.gov/23077180/.

    Article  CAS  Google Scholar 

  42. Khursheed M, Kolla JN, Kotapalli V, Gupta N, Gowrishankar S, Uppin SG, et al. ARID1B, a member of the human SWI/SNF chromatin remodeling complex, exhibits tumour-suppressor activities in pancreatic cancer cell lines. Br J Cancer. 2013;108(10):2056–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Tordella L, Khan S, Hohmeyer A, Banito A, Klotz S, Raguz S, et al. SWI/SNF regulates a transcriptional program that induces senescence to prevent liver cancer. Genes Dev. 2016;30(19):2187–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Sarnowska E, Gratkowska DM, Sacharowski SP, Cwiek P, Tohge T, Fernie AR, et al. The Role of SWI/SNF Chromatin Remodeling Complexes in Hormone Crosstalk. Trends Plant Sci. 2016;21(7):594–608.

    Article  CAS  PubMed  Google Scholar 

  45. Kwon SM, Hong SM, Lee YK, Min S, Yoon G. Metabolic features and regulation in cell senescence. BMB Rep. 2019;52(1):5–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Nacarelli T, Sell C. Targeting metabolism in cellular senescence, a role for intervention. Mol Cell Endocrinol. 2017;455:83–92.

    Article  CAS  PubMed  Google Scholar 

  47. Birsoy K, Wang T, Chen WW, Freinkman E, Abu-Remaileh M, Sabatini DM. An Essential Role of the Mitochondrial Electron Transport Chain in Cell Proliferation Is to Enable Aspartate Synthesis. Cell. 2015;162(3):540–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Sullivan LB, Gui DY, Hosios AM, Bush LN, Freinkman E, Vander Heiden MG. Supporting Aspartate Biosynthesis Is an Essential Function of Respiration in Proliferating Cells. Cell. 2015;162(3):552–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Aird KM, Zhang R. Nucleotide metabolism, oncogene-induced senescence and cancer. Cancer Lett. 2015;356(2 Pt A):204–10.

    Article  CAS  PubMed  Google Scholar 

  50. McHugh D, Gil J. Senescence and aging: Causes, consequences, and therapeutic avenues. J Cell Biol. 2018;217(1):65–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Acosta JC, O’Loghlen A, Banito A, Guijarro MV, Augert A, Raguz S, et al. Chemokine Signaling via the CXCR2 Receptor Reinforces Senescence. Cell. 2008;133(6):1006–18.

    Article  CAS  PubMed  Google Scholar 

  52. 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(8):978–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kuilman T, Michaloglou C, Vredeveld LCW, Douma S, van Doorn R, Desmet CJ, et al. Oncogene-Induced Senescence Relayed by an Interleukin-Dependent Inflammatory Network. Cell. 2008;133(6):1019–31.

    Article  CAS  PubMed  Google Scholar 

  54. Kang TW, Yevsa T, Woller N, Hoenicke L, Wuestefeld T, Dauch D, et al. Senescence surveillance of pre-malignant hepatocytes limits liver cancer development. Nature. 2011;479(7374):547–51.

    Article  CAS  PubMed  Google Scholar 

  55. Coppe JP, Kauser K, Campisi J, Beauséjour CM. Secretion of vascular endothelial growth factor by primary human fibroblasts at senescence. J Biol Chem. 2006;281(40):29568–74.

    Article  CAS  PubMed  Google Scholar 

  56. 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(7525):134–7.

    Article  PubMed  CAS  Google Scholar 

  57. 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(4):533–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Gorgoulis V, Adams PD, Alimonti A, Bennett DC, Bischof O, Bishop C, et al. Cellular Senescence: Defining a Path Forward. Cell. 2019;179(4):813–27.

    Article  CAS  PubMed  Google Scholar 

  59. Carroll B, Korolchuk VI. Dysregulation of mTORC1/autophagy axis in senescence. Aging. 2017;9(8):1851–2.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Narita M, Young ARJ, Arakawa S, Samarajiwa SA, Nakashima T, Yoshida S, et al. Spatial coupling of mTOR and autophagy augments secretory phenotypes. Science. 2011;332(6032):966–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Brusuker I, Rhodes JM, Goldman R. β-Galactosidase—An indicator of the maturational stage of mouse and human mononuclear phagocytes. J Cell Physiol. 1982;112(3):385–90.

    Article  Google Scholar 

  62. Kopp HG, Hooper AT, Shmelkov SV, Rafii S. β-galactosidase staining on bone marrow. The osteoclast pitfall. Histol Histopathol. 2007;22(9):971–6.

    CAS  PubMed  Google Scholar 

  63. Dimri GP, Lee X, Basile G, Acosta M, Scott G, Roskelley C, et al. A biomarker that identifies senescent human cells in culture and in aging skin in vivo. Proc Natl Acad Sci U S A. 1995;92(20):9363–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Muñoz-Espín D, Cañamero M, Maraver A, Gómez-López G, Contreras J, Murillo-Cuesta S, et al. Programmed cell senescence during mammalian embryonic development. Cell. 2013;155(5):1104–18.

    Article  PubMed  CAS  Google Scholar 

  65. Storer M, Mas A, Robert-Moreno A, Pecoraro M, Ortells MC, Di Giacomo V, et al. Senescence is a developmental mechanism that contributes to embryonic growth and patterning. Cell. 2013;155(5):1119–30.

    Article  CAS  PubMed  Google Scholar 

  66. Chapman J, Fielder E, Passos JF. Mitochondrial dysfunction and cell senescence: deciphering a complex relationship. FEBS Lett. 2019;593(13):1566–79.

    Article  CAS  PubMed  Google Scholar 

  67. Dasari A, Bartholomew JN, Volonte D, Galbiati F. Oxidative stress induces premature senescence by stimulating caveolin-1 gene transcription through p38 mitogen-activated protein kinase/Sp1-mediated activation of two GC-rich promoter elements. Cancer Res. 2006;66(22):10805–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Freund A, Laberge RM, Demaria M, Campisi J. Lamin B1 loss is a senescence-associated biomarker. Mol Biol Cell. 2012;23(11):2066–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Wang AS, Ong PF, Chojnowski A, Clavel C, Dreesen O. Loss of lamin B1 is a biomarker to quantify cellular senescence in photoaged skin. Sci Rep. 2017;7(1):15678.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Hayflick L, Moorhead PS. The serial cultivation of human diploid cell strains. Exp Cell Res. 1961;25(3):585–621.

    Article  CAS  PubMed  Google Scholar 

  71. de Magalhães JP, Passos JF. Stress, cell senescence and organismal ageing. Mech Ageing Dev. 2018;170:2–9.

    Article  PubMed  CAS  Google Scholar 

  72. Blasco MA. Telomeres and human disease: Ageing, cancer and beyond. Nat Rev Genet. 2005;6(8):611–22.

    Article  CAS  PubMed  Google Scholar 

  73. Martínez P, Blasco MA. Telomere-driven diseases and telomere-targeting therapies. J Cell Biol. 2017;216(4):875–87.

  74. Olovnikov AM. A theory of marginotomy. The incomplete copying of template margin in enzymic synthesis of polynucleotides and biological significance of the phenomenon. J Theor Biol. 1973;41(1):181–90.

    Article  CAS  PubMed  Google Scholar 

  75. Harley CB, Futcher AB, Greider CW. Telomeres shorten during ageing of human fibroblasts. Nature. 1990;345(6274):458–60.

    Article  CAS  PubMed  Google Scholar 

  76. Greider CW, Blackburn EH. Identification of a specific telomere terminal transferase activity in tetrahymena extracts. Cell. 1985;43(2 Pt 1):405–13.

    Article  CAS  PubMed  Google Scholar 

  77. Bodnar AG, Ouellette M, Frolkis M, Holt SE, Chiu CP, Morin GB, et al. Science. 1998;279(5349):349–52.

    Article  CAS  PubMed  Google Scholar 

  78. Blasco MA, Lee HW, Hande MP, Samper E, Lansdorp PM, RA DP, et al. Telomere shortening and tumor formation by mouse cells lacking telomerase RNA. Cell. 1997;91(1):25–34.

    Article  CAS  PubMed  Google Scholar 

  79. Rudolph KL, Chang S, Lee HW, Blasco M, Gottlieb GJ, Greider C, et al. Longevity, stress response, and cancer in aging telomerase-deficient mice. Cell. 1999;96(5):701–12.

    Article  CAS  PubMed  Google Scholar 

  80. Jaskelioff M, Muller FL, Paik JH, Thomas E, Jiang S, Adams AC, et al. Telomerase reactivation reverses tissue degeneration in aged telomerase-deficient mice. Nature. 2011;469(7328):102–6.

    Article  CAS  PubMed  Google Scholar 

  81. Campisi J. Aging, Cellular Senescence, and Cancer. Annu Rev Physiol. 2013;75(1):685–705.

    Article  CAS  PubMed  Google Scholar 

  82. Itahana K, Campisi J, Dimri GP. Mechanisms of cellular senescence in human and mouse cells. Biogerontology. 2004;5(1):1–10.

    Article  CAS  PubMed  Google Scholar 

  83. Calado RT, Dumitriu B. Telomere dynamics in mice and humans. Semin Hematol. 2013;50(2):165–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Sharpless NE, Sherr CJ. Forging a signature of in vivo senescence. Nat Rev Cancer. 2015;15(7):397–408.

    Article  CAS  PubMed  Google Scholar 

  85. Rodier F, Campisi J. Four faces of cellular senescence. J Cell Biol. 2011;192(4):547–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. D’Adda Di Fagagna F. Living on a break: Cellular senescence as a DNA-damage response. Nat Rev Cancer. 2008;8(7):512–22.

    Article  PubMed  CAS  Google Scholar 

  87. Zhang Y, Xiong Y, Yarbrough WG. ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell. 1998;92(6):725–34.

    Article  CAS  PubMed  Google Scholar 

  88. Rufini A, Tucci P, Celardo I, Melino G. Senescence and aging: The critical roles of p53. Oncogene. 2013;32(43):5129–43.

    Article  CAS  PubMed  Google Scholar 

  89. Pantoja C, Serrano M. Murine fibroblasts lacking p21 undergo senescence and are resistant to transformation by oncogenic Ras. Oncogene. 1999;18(35):4974–82.

    Article  CAS  PubMed  Google Scholar 

  90. Martín-Caballero J, Serrano M, Flores JM, García-Palencia P. Tumor susceptibility of p21waf1/cip1-deficient mice. Cancer Res. 2001;61(16):6234–8.

    PubMed  Google Scholar 

  91. Quereda V, Porlan E, Canãmero M, Dubus P, Malumbres M. An essential role for Ink4 and Cip/Kip cell-cycle inhibitors in preventing replicative stress. Cell Death Differ. 2016;23(3):430–41.

    Article  CAS  PubMed  Google Scholar 

  92. Kim WY, Sharpless NE. The Regulation of INK4/ARF in Cancer and Aging. Cell. 2006;127(2):265–75.

    Article  CAS  PubMed  Google Scholar 

  93. Classon M, Harlow E. The retinoblastoma tumour suppressor in development and cancer. Nat Rev Cancer. 2002;2(12):910–7.

    Article  CAS  PubMed  Google Scholar 

  94. Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, Donovan J, et al. The landscape of somatic copy-number alteration across human cancers. Nature. 2010;463(7283):899–905.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Esteller M, Corn PG, Baylin SB, Herman JG. A gene hypermethylation profile of human cancer. Cancer Res. 2001;61(8):3225–9.

    CAS  PubMed  Google Scholar 

  96. LaPak KM, Burd CE. The molecular balancing act of p16ink4a in cancer and aging. Mol Cancer Res. 2014;12(2):167–83.

    Article  CAS  PubMed  Google Scholar 

  97. Serrano M, Lin AW, McCurrach ME, Beach D, Lowe SW. Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16(INK4a). Cell. 1997;88(5):593–602.

    Article  CAS  PubMed  Google Scholar 

  98. Lin AW, Barradas M, Stone JC, Van Aelst L, Serrano M, Lowe SW. Premature senescence involving p53 and p16 is activated in response to constitutive MEK/MAPK mitogenic signaling. Genes Dev. 1998;12(19):3008–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, Luise C, et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature. 2006;444(7119):638–42.

    Article  PubMed  CAS  Google Scholar 

  100. Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, Issaeva N, et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature. 2006;444(7119):633–7.

    Article  CAS  PubMed  Google Scholar 

  101. Halazonetis TD, Gorgoulis VG, Bartek J. An oncogene-induced DNA damage model for cancer development. Science. 2008;319(5868):1352–5.

    Article  CAS  PubMed  Google Scholar 

  102. Macheret M, Halazonetis TD. Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature. 2018;555(7694):112–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Michaloglou C, Vredeveld LCW, Soengas MS, Denoyelle C, Kuilman T, Van Der Horst CMAM, et al. BRAFE600-associated senescence-like cell cycle arrest of human naevi. Nature. 2005;436(7051):720–4.

    Article  CAS  PubMed  Google Scholar 

  104. Gray-Schopfer VC, Cheong SC, Chong H, Chow J, Moss T, Abdel-Malek ZA, et al. Cellular senescence in naevi and immortalisation in melanoma: A role for p16? Br J Cancer. 2006;95(4):496–505.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Dankort D, Curley DP, Cartlidge RA, Nelson B, Karnezis AN, Damsky WE, et al. BrafV600E cooperates with Pten loss to induce metastatic melanoma. Nat Genet. 2009;41(5):544–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Dhomen N, Reis-Filho JS, da Rocha DS, Hayward R, Savage K, Delmas V, et al. Oncogenic Braf Induces Melanocyte Senescence and Melanoma in Mice. Cancer Cell. 2009;15(4):294–303.

    Article  CAS  PubMed  Google Scholar 

  107. Collado M, Gil J, Efeyan A, Guerra C, Schuhmacher AJ, Barradas M, et al. Senescence in premalignant tumours. Nature. 2005;436(7051):642.

    Article  CAS  PubMed  Google Scholar 

  108. Chen Z, Trotman LC, Shaffer D, Lin HK, Dotan ZA, Niki M, et al. Crucial role of p53-dependent cellular senescence in suppression of Pten-deficient tumorigenesis. Nature. 2005;436(7051):725–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Braig M, Lee S, Loddenkemper C, Rudolph C, AHFM P, Schlegelberger B, et al. Oncogene-induced senescence as an initial barrier in lymphoma development. Nature. 2005;436(7051):660–5.

    Article  CAS  PubMed  Google Scholar 

  110. Denchi EL, Attwooll C, Pasini D, Helin K. Deregulated E2F Activity Induces Hyperplasia and Senescence-Like Features in the Mouse Pituitary Gland. Mol Cell Biol. 2005;25(7):2660–72.

    Article  CAS  PubMed Central  Google Scholar 

  111. Mooi WJ. Oncogene-induced cellular senescence: Causal factor in the growth arrest of pituitary microadenomas? Horm Res. 2009;71(Suppl 2):78–81.

    CAS  PubMed  Google Scholar 

  112. Calcinotto A, Kohli J, Zagato E, Pellegrini L, Demaria M, Alimonti A. Cellular senescence: Aging, cancer, and injury. Physiol Rev. 2019;99(2):1047–78.

    Article  CAS  PubMed  Google Scholar 

  113. Mikuła-Pietrasik J, Niklas A, Uruski P, Tykarski A, Książek K. Mechanisms and significance of therapy-induced and spontaneous senescence of cancer cells. Cell Mol Life Sci. 2020;77(2):213–29.

    Article  PubMed  CAS  Google Scholar 

  114. Chang BD, Broude EV, Dokmanovic M, Zhu H, Ruth A, Xuan Y, et al. A senescence-like phenotype distinguishes tumor cells that undergo terminal proliferation arrest after exposure to anticancer agents. Cancer Res. 1999;59(15):3761–7.

    CAS  PubMed  Google Scholar 

  115. Schmitt CA, Fridman JS, Yang M, Lee S, Baranov E, Hoffman RM, et al. A senescence program controlled by p53 and p16INK4a contributes to the outcome of cancer therapy. Cell. 2002;109(3):335–46.

    Article  CAS  PubMed  Google Scholar 

  116. te Poele RH, Okorokov AL, Jardine L, Cummings J, Joel SP. DNA damage is able to induce senescence in tumor cells in vitro and in vivo. Cancer Res. 2002;62(6):1876–83.

    Google Scholar 

  117. Roninson IB. Tumor cell senescence in cancer treatment. Cancer Res. 2003;63(11):2705–15.

    CAS  PubMed  Google Scholar 

  118. Jones KR, Elmore LW, Jackson-Cook C, Demasters G, Povirk LF, Holt SE, et al. p53-dependent accelerated senescence induced by ionizing radiation in breast tumour cells. Int J Radiat Biol. 2005;81(6):445–58.

    Article  CAS  PubMed  Google Scholar 

  119. Mirzayans R, Scott A, Cameron M, Murray D. Induction of Accelerated Senescence by γ Radiation in Human Solid Tumor-Derived Cell Lines Expressing Wild-Type TP53. Radiat Res. 2005;163(1):53–62.

    Article  CAS  PubMed  Google Scholar 

  120. Li M, You L, Xue J, Lu Y. Ionizing radiation-induced cellular senescence in normal, non-transformed cells and the involved DNA damage response: A mini review. Front Pharmacol. 2018;9:522.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Shao L, Wang Y, Chang J, Luo Y, Meng A, Zhou D. Hematopoietic stem cell senescence and cancer therapy-induced long-term bone marrow injury. Transl Cancer Res. 2013;2(5):397–411.

    CAS  PubMed  Google Scholar 

  122. Di X, Bright AT, Bellott R, Gaskins E, Robert J, Holt S, et al. A chemotherapy-associated senescence bystander effect in breast cancer cells. Cancer Biol Ther. 2008;7(6):864–72.

    Article  CAS  PubMed  Google Scholar 

  123. 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 

  124. Ameri K, Maltepe E. HIGD1A-mediated dormancy and tumor survival. Mol Cell Oncol. 2015;2(4):e1030537.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  125. Axelson H, Fredlund E, Ovenberger M, Landberg G, Påhlman S. Hypoxia-induced dedifferentiation of tumor cells - A mechanism behind heterogeneity and aggressiveness of solid tumors. Semin Cell Dev Biol. 2005;16(4-5):554–63.

    Article  CAS  PubMed  Google Scholar 

  126. Kobayashi A, Okuda H, Xing F, Pandey PR, Watabe M, Hirota S, et al. Bone morphogenetic protein 7 in dormancy and metastasis of prostate cancer stem-like cells in bone. J Exp Med. 2011;208(13):2641–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Wang N, Docherty F, Brown HK, Reeves K, Fowles A, Lawson M, et al. Mitotic quiescence, but not unique “stemness,” marks the phenotype of bone metastasis-initiating cells in prostate cancer. FASEB J. 2015;29(8):3141–50.

    Article  CAS  PubMed  Google Scholar 

  128. Hofstetter CP, Burkhardt JK, Shin BJ, Gürsel DB, Mubita L, Gorrepati R, et al. Protein phosphatase 2A mediates dormancy of glioblastoma multiforme-derived tumor stem-like cells during hypoxia. PLoS One. 2012;7(1):e30059.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Harman D. Aging: a theory based on free radical and radiation chemistry. J Gerontol. 1956;11(3):298–300.

    Article  CAS  PubMed  Google Scholar 

  130. Sies H. Oxidative stress: A concept in redox biology and medicine. Redox Biol. 2015;4:180–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Höhn A, Weber D, Jung T, Ott C, Hugo M, Kochlik B, et al. Happily (n) ever after: Aging in the context of oxidative stress, proteostasis loss and cellular senescence. Redox Biol [Internet]. 2017;11(November 2016):482–501 Available from: https://doi.org/10.1016/j.redox.2016.12.001.

    Article  PubMed  CAS  Google Scholar 

  132. Cadet J, Richard WJ. DNA base damage by reactive oxygen species, oxidizing agents, and UV radiation. Cold Spring Harb Perspect Biol. 2013;5(2):a012559.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  133. Von Zglinicki T. Oxidative stress shortens telomeres. Trends Biochem Sci. 2002;27(7):339–44.

    Article  Google Scholar 

  134. Passos JF, Nelson G, Wang C, Richter T, Simillion C, Proctor CJ, et al. Feedback between p21 and reactive oxygen production is necessary for cell senescence. Mol Syst Biol. 2010;6:347.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  135. Chen Y, Liu K, Shi Y, Shao C. The tango of ROS and p53 in tissue stem cells. Cell Death Differ. 2018;25(4):637–9.

    Article  PubMed Central  Google Scholar 

  136. Höhn A, Jung T, Grune T. Pathophysiological importance of aggregated damaged proteins. Free Radic Biol Med. 2014;71:70–89.

    Article  PubMed  CAS  Google Scholar 

  137. Zhang YG, Wang L, Kaifu T, Li J, Li X, Li L. Featured Article: Accelerated decline of physical strength in peroxiredoxin-3 knockout mice. Exp Biol Med. 2016;241(13):1395–400.

    Article  CAS  Google Scholar 

  138. Nyström T, Yang J, Molin M. Peroxiredoxins, gerontogenes linking aging to genome instability and cancer. Genes Dev. 2012;26(18):2001–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Picot CR, Perichon M, Cintrat JC, Friguet B, Petropoulos I. The peptide methionine sulfoxide reductases, MsrA and MsrB (hCBS-1), are downregulated during replicative senescence of human WI-38 fibroblasts. FEBS Lett. 2004;558(1-3):74–8.

    Article  CAS  PubMed  Google Scholar 

  140. Liao N, Shi Y, Zhang C, Zheng Y, Wang Y, Zhao B, et al. Antioxidants inhibit cell senescence and preserve stemness of adipose tissue-derived stem cells by reducing ROS generation during long-term in vitro expansion. Stem Cell Res Ther. 2019;10(1):306.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  141. Kornienko JS, Smirnova IS, Pugovkina NA, Ivanova JS, Shilina MA, Grinchuk TM, et al. High doses of synthetic antioxidants induce premature senescence in cultivated mesenchymal stem cells. Sci Rep. 2019;9(1):1296.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  142. López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G. The hallmarks of aging. Cell. 2013;153(6):1194.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  143. Pérez VI, Van Remmen H, Bokov A, Epstein CJ, Vijg J, Richardson A. The overexpression of major antioxidant enzymes does not extend the lifespan of mice. Aging Cell. 2009;1790(10):1005–14.

    Google Scholar 

  144. Dolivo D, Hernandez S, Dominko T. Cellular lifespan and senescence: a complex balance between multiple cellular pathways. Insid Cell. 2016;1(1):36–47.

    Article  CAS  Google Scholar 

  145. Hekimi S, Lapointe J, Wen Y. Taking a “good” look at free radicals in the aging process. Trends Cell Biol. 2011;21(10):569–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Yun J, Finkel T. Mitohormesis. Cell Metab. 2014;19(5):757–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Vaupel P, Kallinowski F, Okunieff P. Blood Flow, Oxygen and Nutrient Supply, and Metabolic Microenvironment of Human Tumors: A Review. Cancer Res. 1989;49(23):6449–65.

    CAS  PubMed  Google Scholar 

  148. Carreau A, Hafny-Rahbi BE, Matejuk A, Grillon C, Kieda C. Why is the partial oxygen pressure of human tissues a crucial parameter? Small molecules and hypoxia. J Cell Mol Med [Internet]. 2011;15(6):1239–53 [cited 2021 Apr 12], Available from: /pmc/articles/PMC4373326/.

    Article  CAS  Google Scholar 

  149. Pouyssegur J, López-Barneo J. Hypoxia in health and disease. Mol Asp Med. 2016;47–48:1–2.

    Article  Google Scholar 

  150. Lopez-Barneo J. Oxygen-sensing by ion channels and the regulation of cellular functions. Trends Neurosci. 1996;19(10):435–40.

    Article  CAS  PubMed  Google Scholar 

  151. Weir EK, López-Barneo J, Buckler KJ, Archer SL. Mechanisms of disease: Acute oxygen-sensing mechanisms. N Engl J Med. 2005;353(19):2042–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. López-Barneo J, Simon MC. Cellular adaptation to oxygen deficiency beyond the Nobel award. Nat Commun. 2020;11(1):607.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  153. Lee JW, Ko J, Ju C, Eltzschig HK. Hypoxia signaling in human diseases and therapeutic targets. Exp Mol Med [Internet]. 2019;51(6):1–13 Available from: https://doi.org/10.1038/s12276-019-0235-1.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. Jewell UR, Kvietikova I, Scheid A, Bauer C, Wenger RH, Gassmann M. Induction of HIF-1alpha in response to hypoxia is instantaneous. FASEB J. 2001;15(7):1312–4.

    Article  CAS  PubMed  Google Scholar 

  155. Fandrey J, Gorr TA, Gassmann M. Regulating cellular oxygen sensing by hydroxylation. Cardiovasc Res. 2006;71(4):642–51.

    Article  CAS  PubMed  Google Scholar 

  156. Majmundar AJ, Wong WJ, Simon MC. Hypoxia-Inducible Factors and the Response to Hypoxic Stress. Mol Cell [Internet]. 2010;40(2):294–309 Available from: https://doi.org/10.1016/j.molcel.2010.09.022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Kaelin WG, Ratcliffe PJ. Oxygen Sensing by Metazoans: The Central Role of the HIF Hydroxylase Pathway. Mol Cell. 2008;30(4):393–402.

    Article  CAS  PubMed  Google Scholar 

  158. Franke K, Gassmann M, Wielockx B. Erythrocytosis: The HIF pathway in control. Blood. 2013;122(7):1122–8.

    Article  CAS  PubMed  Google Scholar 

  159. Gu YZ, Moran SM, Hogenesch JB, Wartman L, Bradfield CA. Molecular characterization and chromosomal localization of a third α- class hypoxia inducible factor subunit, HIF3α. Gene Expr. 1998;7(3):205–13.

    CAS  PubMed  Google Scholar 

  160. Packer L, Fuehr K. Low oxygen concentration extends the lifespan of cultured human diploid cells [13]. Nature. 1977;267(5610):423–5.

    Article  CAS  PubMed  Google Scholar 

  161. Saito H, Hammond AT, Moses RE. The effect of low oxygen tension on the in vitro-replicative life span of human diploid fibroblast cells and their transformed derivatives. Exp Cell Res. 1995;217(2):272–9.

    Article  CAS  PubMed  Google Scholar 

  162. 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(8):741–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Betts DH, Perrault SD, King WA. Low oxygen delays fibroblast senescence despite shorter telomeres. Biogerontology. 2008;9(1):19–31.

    Article  CAS  PubMed  Google Scholar 

  164. Kamijo T, Zindy F, Roussel MF, Quelle DE, Downing JR, Ashmun RA, et al. Tumor suppression at the mouse INK4a locus mediated by the alternative reading frame product p19(ARF). Cell. 1997;91(5):649–59.

    Article  CAS  PubMed  Google Scholar 

  165. Wright WE, Shay JW. Telomere dynamics in cancer progression and prevention: Fundamental differences in human and mouse telomere biology. Nat Med. 2000;6(8):849–51.

    Article  CAS  PubMed  Google Scholar 

  166. Patrick A, Seluanov M, Hwang C, Tam J, Khan T, Morgenstern A, et al. Sensitivity of primary fibroblasts in culture to atmospheric oxygen does not correlate with species lifespan. Aging (Albany NY) [Internet]. 2016;8(5):841–7 [cited 2021 Mar 25], Available from: /pmc/articles/PMC4931838/.

    CAS  Google Scholar 

  167. Kondoh H, Lleonart ME, Gil J, Wang J, Degan P, Peters G, et al. Glycolytic enzymes can modulate cellular life span. Cancer Res. 2005;65(1):177–85.

    Article  CAS  PubMed  Google Scholar 

  168. Tsai CC, Chen YJ, Yew TL, Chen LL, Wang JY, Chiu CH, et al. Hypoxia inhibits senescence and maintains mesenchymal stem cell properties through down-regulation of E2A-p21 by HIF-TWIST. Blood. 2011;117(2):459–69.

    Article  CAS  PubMed  Google Scholar 

  169. Gordan JD, Bertout JA, Hu CJ, Diehl JA, Simon MC. HIF-2α Promotes Hypoxic Cell Proliferation by Enhancing c-Myc Transcriptional Activity. Cancer Cell. 2007;11(4):335–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Bell EL, Klimova TA, Eisenbart J, Schumacker PT, Chandel NS. Mitochondrial Reactive Oxygen Species Trigger Hypoxia-Inducible Factor-Dependent Extension of the Replicative Life Span during Hypoxia. Mol Cell Biol. 2007;27(16):5737–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Sermeus A, Michiels C. Reciprocal influence of the p53 and the hypoxic pathways. Cell Death Dis. 2011;2(5):e164.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Leontieva OV, Natarajan V, Demidenko ZN, Burdelya LG, Gudkov AV, Blagosklonny MV. Hypoxia suppresses conversion from proliferative arrest to cellular senescence. Proc Natl Acad Sci U S A. 2012;109(33):13314–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Koshiji M, Kageyama Y, Pete EA, Horikawa I, Barrett JC, Huang LE. HIF-1α induces cell cycle arrest by functionally counteracting Myc. EMBO J. 2004;23(9):1949–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Gardner LB, Li Q, Park MS, Flanagan WM, Semenza GL, Dang CV. Hypoxia Inhibits G1/S Transition through Regulation of p27 Expression. J Biol Chem. 2001;276(11):7919–26.

    Article  CAS  PubMed  Google Scholar 

  175. Goda N, Ryan HE, Khadivi B, McNulty W, Rickert RC, Johnson RS. Hypoxia-Inducible Factor 1 Is Essential for Cell Cycle Arrest during Hypoxia. Mol Cell Biol. 2003;23(1):359–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Hammer S, To KKW, Yoo YG, Koshiji M, Huang LE. Hypoxic suppression of the cell cycle gene CDC25A in tumor cells. Cell Cycle. 2007;6(15):1919–26.

    Article  CAS  PubMed  Google Scholar 

  177. Welford SM, Giaccia AJ. Hypoxia and senescence: The impact of oxygenation on tumor suppression. Mol Cancer Res. 2011;9(5):538–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Young AP, Schisio S, Minamishima YA, Zhang Q, Li L, Grisanzio C, et al. VHL loss actuates a HIF-independent senescence programme mediated by Rb and p400. Nat Cell Biol. 2008;10(3):361–9.

    Article  CAS  PubMed  Google Scholar 

  179. Wang W, Wang D, Li H. Initiation of premature senescence by Bcl-2 in hypoxic condition. Int J Clin Exp Pathol. 2014;7(5):2446–53.

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Webster WS, Abela D. The effect of hypoxia in development. Birth Defects Res C Embryo Today Rev. 2007;81(3):215–28.

    Article  CAS  Google Scholar 

  181. Podkalicka P, Stępniewski J, Mucha O, Kachamakova-Trojanowska N, Dulak J, Łoboda A. Hypoxia as a Driving Force of Pluripotent Stem Cell Reprogramming and Differentiation to Endothelial Cells. Biomolecules [Internet]. 2020;10(12):1614 [cited 2021 Mar 25]Available from: https://www.mdpi.com/2218-273X/10/12/1614.

    Article  CAS  Google Scholar 

  182. Simon MC, Keith B. The role of oxygen availability in embryonic development and stem cell function [Internet]. Nat Rev Mol Cell Biol Nature Publishing Group. 2008;9:285–96 [cited 2021 Mar 25], Available from: www.nature.com/reviews/molcellbio.

    Article  CAS  Google Scholar 

  183. Cowden Dahl KD, Fryer BH, Mack FA, Compernolle V, Maltepe E, Adelman DM, et al. Hypoxia-Inducible Factors 1α and 2α Regulate Trophoblast Differentiation. Mol Cell Biol [Internet]. 2005;25(23):10479–91 [cited 2021 Mar 25], Available from: https://pubmed.ncbi.nlm.nih.gov/16287860/.

    Article  CAS  Google Scholar 

  184. Mohyeldin A, Garzón-Muvdi T, Quiñones-Hinojosa A. Oxygen in stem cell biology: A critical component of the stem cell niche. Cell Stem Cell. 2010;7(2):150–61.

    Article  CAS  PubMed  Google Scholar 

  185. Huang G, Ye S, Zhou X, Liu D, Ying QL. Molecular basis of embryonic stem cell self-renewal: From signaling pathways to pluripotency network. Cell Mol Life Sci. 2015;72(9):1741–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Danet GH, Pan Y, Luongo JL, Bonnet DA, Simon MC. Expansion of human SCID-repopulating cells under hypoxic conditions. J Clin Invest. 2003;112(1):126–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Ezashi T, Das P, Roberts RM. Low O2 tensions and the prevention of differentiation of hES cells. Proc Natl Acad Sci U S A. 2005;102(13):4783–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Morrison SJ, Csete M, Groves AK, Melega W, Wold B, Anderson DJ. Culture in reduced levels of oxygen promotes clonogenic sympathoadrenal differentiation by isolated neural crest stem cells. J Neurosci. 2000;20(19):7370–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Nugud A, Sandeep D, El-Serafi AT. Two faces of the coin: Minireview for dissecting the role of reactive oxygen species in stem cell potency and lineage commitment [Internet]. J Adv Res Elsevier BV. 2018;14:73–9 [cited 2021 Mar 25], Available from: https://pubmed.ncbi.nlm.nih.gov/30023134/.

    Article  CAS  Google Scholar 

  190. Takahashi K, Yamanaka S. Induction of Pluripotent Stem Cells from Mouse Embryonic and Adult Fibroblast Cultures by Defined Factors. Cell. 2006;126(4):663–76.

    Article  CAS  PubMed  Google Scholar 

  191. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, et al. Induction of Pluripotent Stem Cells from Adult Human Fibroblasts by Defined Factors. Cell. 2007;131(5):861–72.

    Article  CAS  PubMed  Google Scholar 

  192. Huangfu D, Maehr R, Guo W, Eijkelenboom A, Snitow M, Chen AE, et al. Induction of pluripotent stem cells by defined factors is greatly improved by small-molecule compounds. Nat Biotechnol. 2008;26(7):795–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Okita K, Hong H, Takahashi K, Yamanaka S. Generation of mouse-induced pluripotent stem cells with plasmid vectors. Nat Protoc. 2010;5(3):418–28.

    Article  CAS  PubMed  Google Scholar 

  194. Sugimoto K, Matsuura T, Nakazono A, Igawa K, Yamada S, Hayashi Y. Effects of hypoxia inducible factors on pluripotency in human iPS cells. Microsc Res Tech. 2018;81(7):749–54.

    Article  CAS  PubMed  Google Scholar 

  195. Mathieu J, Zhou W, Xing Y, Sperber H, Ferreccio A, Agoston Z, et al. Hypoxia-inducible factors have distinct and stage-specific roles during reprogramming of human cells to pluripotency. Cell Stem Cell. 2014;14(5):592–605.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Prigione A, Rohwer N, Hoffmann S, Mlody B, Drews K, Bukowiecki R, et al. HIF1α modulates cell fate reprogramming through early glycolytic shift and upregulation of PDK1-3 and PKM2. Stem Cells. 2014;32(2):364–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Banito A, Rashid ST, Acosta JC, De Li S, Pereira CF, Geti I, et al. Senescence impairs successful reprogramming to pluripotent stem cells. Genes Dev. 2009;23(18):2134–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Li H, Collado M, Villasante A, Strati K, Ortega S, Cãamero M, et al. The Ink4/Arf locus is a barrier for iPS cell reprogramming. Nature. 2009;460(7259):1136–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Mosteiro L, Pantoja C, Alcazar N, Marión RM, Chondronasiou D, Rovira M, et al. Tissue damage and senescence provide critical signals for cellular reprogramming in vivo. Science (80- ). 2016;354(6315):aaf4445.

    Article  CAS  Google Scholar 

  200. Mosteiro L, Pantoja C, de Martino A, Serrano M. Senescence promotes in vivo reprogramming through p16 INK4a and IL-6. Aging Cell. 2018;17(2):e12711.

    Article  CAS  Google Scholar 

  201. Chiche A, Le Roux I, von Joest M, Sakai H, Aguín SB, Cazin C, et al. Injury-Induced Senescence Enables In Vivo Reprogramming in Skeletal Muscle. Cell Stem Cell. 2017;20(3):407–414.e4.

    Article  CAS  PubMed  Google Scholar 

  202. Ocampo A, Reddy P, Martinez-Redondo P, Platero-Luengo A, Hatanaka F, Hishida T, et al. In Vivo Amelioration of Age-Associated Hallmarks by Partial Reprogramming. Cell. 2016;167(7):1719–1733.e12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Milanovic M, Fan DNY, Belenki D, Däbritz JHM, Zhao Z, Yu Y, et al. Senescence-associated reprogramming promotes cancer stemness. Nature [Internet]. 2018;553(7686):96–100 Available from: https://doi.org/10.1038/nature25167.

    Article  CAS  PubMed  Google Scholar 

  204. 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. Genes Dev. 2017;31(2):172–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Hanahan D, Weinberg RA. Hallmarks of cancer: The next generation. Cell. 2011;144(5):646–74.

    Article  CAS  PubMed  Google Scholar 

  206. Al-Hajj M, Wicha MS, Benito-Hernandez A, Morrison SJ, Clarke MF. Prospective identification of tumorigenic breast cancer cells. Proc Natl Acad Sci U S A. 2003;100(7):3983–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Kalluri R, Zeisberg M. Fibroblasts in cancer. Nat Rev Cancer. 2006;6(5):392–401.

    Article  CAS  PubMed  Google Scholar 

  208. Junttila MR, De Sauvage FJ. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature. 2013;501(7467):346–54.

    Article  CAS  PubMed  Google Scholar 

  209. Brurberg KG, Thuen M, Ruud E-BM, Rofstad EK. Fluctuations in p O 2 in Irradiated Human Melanoma Xenografts. Radiat Res. 2006;165(1):16–25.

    Article  CAS  PubMed  Google Scholar 

  210. Bennewith KL, Durand RE. Quantifying transient hypoxia in human tumor xenografts by flow cytometry. Cancer Res. 2004;64(17):6183–9.

    Article  CAS  PubMed  Google Scholar 

  211. Azuma T, Yao S, Zhu G, Flies AS, Flies SJ, Chen L. B7-H1 is a ubiquitous antiapoptotic receptor on cancer cells. Blood. 2008;111(7):3635–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Triana-Martínez F, Loza MI, Domínguez E. Beyond Tumor Suppression: Senescence in Cancer Stemness and Tumor Dormancy. Cells. 2020;9(2):346.

    Article  PubMed Central  CAS  Google Scholar 

  213. Sharma S, Xing F, Liu Y, Wu K, Said N, Pochampally R, et al. Secreted protein acidic and rich in cysteine (sparc) mediates metastatic dormancy of prostate cancer in bone. J Biol Chem [Internet]. 2016;291(37):19351–63 [cited 2021 Jun 21], Available from: https://pubmed.ncbi.nlm.nih.gov/27422817/.

    Article  CAS  Google Scholar 

  214. van Vliet T, Varela-Eirin M, Wang B, Borghesan M, Brandenburg SM, Franzin R, et al. Physiological hypoxia restrains the senescence-associated secretory phenotype via AMPK-mediated mTOR suppression. Mol Cell. 2021;81(9):2041–52.

    Article  PubMed  CAS  Google Scholar 

  215. Conley SJ, Gheordunescu E, Kakarala P, Newman B, Korkaya H, Heath AN, et al. Antiangiogenic agents increase breast cancer stem cells via the generation of tumor hypoxia. Proc Natl Acad Sci U S A. 2012;109(8):2784–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Iriondo O, Rábano M, Domenici G, Carlevaris O, López-Ruiz JA, Zabalza I, et al. Distinct breast cancer stem/progenitor cell populations require either HIF1α or loss of PHD3 to expand under hypoxic conditions. Oncotarget. 2015;6(31):31721–39.

    Article  PubMed  PubMed Central  Google Scholar 

  217. Bhagat M, Palanichamy JK, Ramalingam P, Mudassir M, Irshad K, Chosdol K, et al. HIF-2α mediates a marked increase in migration and stemness characteristics in a subset of glioma cells under hypoxia by activating an Oct-4/Sox-2-Mena (INV) axis. Int J Biochem Cell Biol. 2016;74:60–71.

    Article  CAS  PubMed  Google Scholar 

  218. Johansson E, Grassi ES, Pantazopoulou V, Tong B, Lindgren D, Berg TJ, et al. CD44 Interacts with HIF-2α to Modulate the Hypoxic Phenotype of Perinecrotic and Perivascular Glioma Cells. Cell Rep. 2017;20(7):1641–53.

    Article  CAS  PubMed  Google Scholar 

  219. Das B, Pal B, Bhuyan R, Li H, Sarma A, Gayan S, et al. MYC Regulates the HIF2a Stemness Pathway via Nanog and Sox2 to Maintain Self-Renewal in Cancer Stem Cells versus Non-Stem Cancer Cells. Cancer Res. 2019;79(16):4015–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Wu L, Yu H, Zhao Y, Zhang C, Wang J, Yue X, et al. HIF-2α mediates hypoxia-induced LIF expression in human colorectal cancer cells. Oncotarget. 2015;6(6):4406–17.

    Article  PubMed  PubMed Central  Google Scholar 

  221. Derynck R, Weinberg RA. EMT and Cancer: More Than Meets the Eye. Dev Cell. 2019;49(3):313–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Krishnamachary B, Zagzag D, Nagasawa H, Rainey K, Okuyama H, Baek JH, et al. Hypoxia-inducible factor-1-dependent repression of E-cadherin in von Hippel-Lindau tumor suppressor-null renal cell carcinoma mediated by TCF3, ZFHX1A, and ZFHX1B. Cancer Res. 2006;66(5):2725–31.

    Article  CAS  PubMed  Google Scholar 

  223. Moreno-Bueno G, Portillo F, Cano A. Transcriptional regulation of cell polarity in EMT and cancer. Oncogene. 2008;27(55):6958–69.

    Article  CAS  PubMed  Google Scholar 

  224. Wang T, Gilkes DM, Takano N, Xiang L, Luo W, Bishop CJ, et al. Hypoxia-inducible factors and RAB22A mediate formation of microvesicles that stimulate breast cancer invasion and metastasis. Proc Natl Acad Sci U S A. 2014;111(31):E3234–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. King HW, Michael MZ, Gleadle JM. Hypoxic enhancement of exosome release by breast cancer cells. BMC Cancer. 2012;12:421.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Easwaran H, Tsai HC, Baylin SB. Cancer Epigenetics: Tumor Heterogeneity, Plasticity of Stem-like States, and Drug Resistance. Mol Cell. 2014;54(5):716–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Warburg O. On the origin of cancer cells. Science. 1956;123(3191):309–14.

    Article  CAS  PubMed  Google Scholar 

  228. Chen C, Pore N, Behrooz A, Ismail-Beigi F, Maity A. Regulation of glut1 mRNA by hypoxia-inducible factor-1: Interaction between H-ras and hypoxia. J Biol Chem. 2001;276(12):9519–25.

    Article  CAS  PubMed  Google Scholar 

  229. Mimura I, Nangaku M, Kanki Y, Tsutsumi S, Inoue T, Kohro T, et al. Dynamic Change of Chromatin Conformation in Response to Hypoxia Enhances the Expression of GLUT3 (SLC2A3) by Cooperative Interaction of Hypoxia-Inducible Factor 1 and KDM3A. Mol Cell Biol. 2012;32(15):3018–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  230. Shibuya K, Okada M, Suzuki S, Seino M, Seino S, Takeda H, et al. Targeting the facilitative glucose transporter GLUT1 inhibits the self-renewal and tumor-initiating capacity of cancer stem cells. Oncotarget. 2015;6(2):651–61.

    Article  PubMed  Google Scholar 

  231. Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, Brudno Y, et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science. 2009;324(5929):930–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Wu MZ, Tsai YP, Yang MH, Huang CH, Chang SY, Chang CC, et al. Interplay between HDAC3 and WDR5 Is Essential for Hypoxia-Induced Epithelial-Mesenchymal Transition. Mol Cell. 2011;43(5):811–22.

    Article  CAS  PubMed  Google Scholar 

  233. Kulshreshtha R, Ferracin M, Wojcik SE, Garzon R, Alder H, Agosto-Perez FJ, et al. A MicroRNA Signature of Hypoxia. Mol Cell Biol. 2007;27(5):1859–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Han M, Liu M, Wang Y, Chen X, Xu J, Sun Y, et al. Antagonism of miR-21 reverses epithelial-mesenchymal transition and cancer stem cell phenotype through AKT/ERK1/2 inactivation by targeting PTEN. PLoS One. 2012;7(6):e39520.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Ferraro A, Kontos CK, Boni T, Bantounas I, Siakouli D, Kosmidou V, et al. Epigenetic regulation of miR-21 in colorectal cancer: ITGB4 as a novel miR-21 target and a three-gene network (miR-21-ITGB44-PCDC4) as predictor of metastatic tumor potential. Epigenetics. 2014;9(1):129–41.

    Article  CAS  PubMed  Google Scholar 

  236. Van Den Beucken T, Koch E, Chu K, Rupaimoole R, Prickaerts P, Adriaens M, et al. Hypoxia promotes stem cell phenotypes and poor prognosis through epigenetic regulation of DICER. Nat Commun. 2014;5:5203.

    Article  PubMed  CAS  Google Scholar 

  237. Zhang C, Zhi WI, Lu H, Samanta D, Chen I, Gabrielson E, et al. Hypoxia-inducible factors regulate pluripotency factor expression by ZNF217-and ALKBH5-mediated modulation of RNA methylation in breast cancer cells. Oncotarget. 2016;7(40):64527–42.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

This work is supported by grants from the Ministerio de Ciencia, Innovación y Universidades (MCIU) Plan Estatal de I+D+I 2018, a la Agencia Estatal de Investigación (AEI) y al Fondo Europeo de Desarrollo Regional (MCIU/AEI/FEDER, UE): RTI2018-097455-B-I00; grant from AEI-MICIU/FEDER (RED2018-102723-T); from CIBER de Cáncer (CB16/12/00275), co-funded by FEDER from Regional Development European Funds (European Union); from Consejeria de Salud (PI-0397-2017) and Consejeria of Economía, Conocimiento, Empresas y Universidad of the Junta de Andalucia (P18-RT-2501). Also especial thanks to the Fundacion AECC and Fundacion BBVA for supporting this work.

Funding

This work is supported by grants from the Ministerio de Ciencia, Innovación y Universidades (MCIU) Plan Estatal de I+D+I 2018, a la Agencia Estatal de Investigación (AEI) y al Fondo Europeo de Desarrollo Regional (MCIU/AEI/FEDER, UE): RTI2018-097455-B-I00; grant from AEI-MICIU/FEDER (RED2018-102723-T); from CIBER de Cáncer (CB16/12/00275), co-funded by FEDER from Regional Development European Funds (European Union); from Consejeria de Salud (PI-0397-2017) and Consejeria of Economía, Conocimiento, Empresas y Universidad of the Junta de Andalucia (P18-RT-2501). Also especial thanks to the Fundacion AECC and Fundacion BBVA for supporting this work.

Author information

Authors and Affiliations

Authors

Contributions

Daniel Otero-Albiol and Amancio Carnero conceived and designed this study, drafted and edited the manuscript. The author(s) read and approved the final manuscript.

Authors’ information

Daniel Otero-Albiol and Amancio Carnero, Instituto de Biomedicina de Sevilla (IBIS), Hospital Universitario Virgen del Rocío, Universidad de Sevilla, Consejo Superior de Investigaciones Científicas; Sevilla, Spain and CIBER de Cancer, IS Carlos III, Madrid, Spain.

Corresponding author: Amancio Carnero

Corresponding author

Correspondence to Amancio Carnero.

Ethics declarations

Ethics approval and consent to participate

Not applicable

Consent for publication

Not applicable

Competing interests

Amancio Carnero is member of the Editorial Board of JECCR. The authors declare that they have no other competing interests

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

Otero-Albiol, D., Carnero, A. Cellular senescence or stemness: hypoxia flips the coin. J Exp Clin Cancer Res 40, 243 (2021). https://doi.org/10.1186/s13046-021-02035-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13046-021-02035-0

Keywords