Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Sep 13.
Published in final edited form as: Int Rev Cell Mol Biol. 2017 Apr 28;334:177–205. doi: 10.1016/bs.ircmb.2017.03.008

Senescence-Associated MicroRNAs

Rachel Munk 1, Amaresh C Panda 1, Ioannis Grammatikakis 1, Myriam Gorospe 1, Kotb Abdelmohsen 1,1
PMCID: PMC8436595  NIHMSID: NIHMS1739215  PMID: 28838538

Abstract

Senescent cells arise as a consequence of cellular damage and can have either a detrimental or advantageous impact on tissues and organs depending on the specific cell type and metabolic state. As senescent cells accumulate in tissues with advancing age, they have been implicated in many age-related declines and diseases. The major facets of senescence include two pathways responsible for establishing and maintaining a senescence program, p53/CDKN1A(p21) and CDKN2A(p16)/RB, as well as the senescence-associated secretory phenotype. Numerous MicroRNAs influence senescence by modulating the abundance of key senescence regulatory proteins, generally by lowering the stability and/or translation of mRNAs that encode such factors. Accordingly, understanding the molecular mechanisms by which MicroRNAs influence senescence will enable diagnostic and therapeutic opportunities directed at senescent cells. Here, we review senescence-associated (SA)-MicroRNAs and discuss their implications in senescence-relevant pathologies.

1. INTRODUCTION

1.1. Senescence

Senescence was first described by Leonard Hayflick 50 years ago as a state in which normal cells divide for a finite number of times before they cease proliferation indefinitely (Hayflick, 1965; Hayflick and Moorhead, 1961), although they remain viable and metabolically active (Gey and Seeger, 2013). Senescence has been studied extensively using primary cells cultured ex vivo until they stop dividing; this process is known as replicative senescence (Hayflick, 1992). Additionally, senescence can be achieved by exposure of primary cells to a range of sublethal damaging agents; this process is known as premature senescence. Knowledge from these two processes has been recently complemented by studies on senescence in vivo and its impact in physiology and pathology.

1.1.1. Replicative and Premature Senescence

Replicative senescence is achieved in primary cells isolated from tissues and placed in culture. Over time, their population doubling time (PDL) gradually increases until they are terminally arrested and cease to proliferate. At the same time, the telomeres that protect chromosome ends and permit DNA polymerase to complete replication become progressively shorter and eventually trigger a DNA damage response (DDR) resulting in growth cessation and senescence (Kuilman et al., 2010). The short, unprotected telomeres activate several downstream effectors including the kinases ataxia telangiectasia mutated (ATM) and ATM-related (ATR), the checkpoint kinases CHK1 and CHK2, and the tumor suppressor and transcription factor p53 (Ben-Porath and Weinberg, 2005; Herbig et al., 2004). In turn, p53 transcriptionally induces expression of the cyclin-dependent kinase (CDK) inhibitor CDKN1A/p21 and the alternate open reading frame protein ARF, enabling the p53/p21 axis that inhibits replication (Ben-Porath and Weinberg, 2005; Campisi, 2005). Replicative senescence is further established by the CDKN2A(p16)/RB axis, comprising of the retinoblastoma protein (RB) activated through the CDK inhibitor p16 (CDKN2A/INK4A) and the related proteins p15/INK4B, p18/INK4C, and p19/INK4D (Ben-Porath and Weinberg, 2005; Campisi, 2005).

Premature senescence, also known as stress-inducible senescence, can be rapidly achieved by exposing cells to stresses such as radiation, oxidants, toxins, chemotherapy, oncoprotein activation, or through tumor-suppressor inactivation. These harmful stimuli trigger senescence by activating stress signals including DDR; the ensuing growth arrest by CDK inhibitors and heterochromatin changes occur without apparent loss of telomere function (Kuilman et al., 2010). Premature senescence can be triggered by activation of oncoproteins such as KRAS(V12) and BRAF(V600E), and also by inactivation of the tumor suppressors phosphatase and tensin homologue (PTEN), von Hipple-Lindau (VHL), or neurofibromatosis (NF) 1 (Gorospe and Abdelmohsen, 2011). While ectopic expression of telomerase can restore cell proliferation during replicative senescence, it does not prevent premature senescence (Wei et al., 1999). The stresses that activate senescence also utilize the p53/p21 and p16/RB pathways, as described in a recent review (Loaiza and Demaria, 2016).

1.1.2. In Vivo Senescence

There are many models of replicative and premature senescence, mainly involving cultured primary cells. However, these senescence models remain limited, as they may not fully recapitulate the physiological consequences of senescence. Increasing numbers of studies have provided evidence that senescence occurs in tissues and organs in vivo and can be detected in physiologic and pathologic situations, with both detrimental and protective consequences.

In tissues from humans, mice, and monkeys, senescent cells have been detected showing the enhanced presence of DDR proteins and senescence-associated β-galactosidase activity, as explained later (Cristofalo et al., 2004; Dimri et al., 1995; Herbig et al., 2006). Senescent cells were found to accumulate with age in renewable tissues like epithelia, stroma, and the hematopoietic system (Choi et al., 2000; Dimri et al., 1995; Jeyapalan et al., 2007; Krishnamurthty et al., 2004; Krizhanovsky et al., 2008). Stem cells in muscle from aged mice express p16 as they switch from reversible quiescence to irreversible senescence, suggesting that the maintenance of quiescence depends upon the effective suppression of senescence pathways in vivo (Sousa-Victor et al., 2014). The reliance on p16 for the quiescence-to-senescence transition may explain, at least in part, the elevated proliferation of progenitors of neurons and islet cells in p16-deficient mice, which show improved neuronal and pancreatic function compared with WT mice (Krishnamurthy et al., 2006; Molofsky et al., 2006). Importantly, mice in which p16-expressing cells are selectively eliminated via caspase activation show delayed appearance of age-related conditions, including cataracts, sarcopenia, and loss of adipose tissue (Baker et al., 2011).

In physiologic contexts, senescence is essential for morphogenesis during development. It also plays a pivotal role during wound repair, as the wound-induced senescent fibroblasts and endothelial cells secrete platelet-derived growth factor AA (PDGF-AA) which accelerates myofibroblast differentiation, wound healing, and tissue repair (Demaria et al., 2014). Similarly, following liver damage, hepatic stellate cells acquire senescence, thereby increasing degradation of the extracellular matrix, preventing the formation of fibrotic scars, and preserving liver function (Krizhanovsky et al., 2008).

On the other hand, senescence has also been detected and implicated in various pathologies, prominently cancer. Interestingly, the same oncogenes that can trigger senescence (e.g., KRAS(V12), BRAF (V600E)) can also trigger tumorigenesis if cells are unable to mount a senescent response (Braig et al., 2005; Collado et al., 2005; Dankort et al., 2009; Michaloglou et al., 2005). The complexity between senescence and cancer is further illustrated by the senescence-associated secretory phenotype (SASP), as this phenotype can favor or suppress tumorigenesis. Favoring tumorigenesis is the secretion by senescent cells of angiogenic factors (e.g., VEGF) that promote local angiogenesis, and matrix metalloproteinases (MMPs) which make tissues permeable and facilitate both migration and metastasis (Freund et al., 2010; Ohtani et al., 2012). Suppressing tumorigenesis, particularly at certain tumor stages, are SASP factors that promote chemotaxis leading to the recognition and clearance of cancer cells by the immune system (Acosta et al., 2008; Kuilman et al., 2008; Massague, 2008).

Senescent cells can also exacerbate other pathologic situations. They have been implicated in age-related diseases such as atherosclerosis, diabetes, neurodegeneration, and cardiovascular illnesses (Childs et al., 2015; Chinta et al., 2015). For example, accumulation of senescent microglia with aging is linked to neurodegenerative diseases such as Parkinson’s and Alzheimer’s diseases (Luo et al., 2010), while the impaired vascularization in diabetes is linked to the enhanced senescence of endothelial progenitor cells (Rosso et al., 2006). Finally, senescence of skeletal muscle and precursor cells is linked to muscle aging and the related disease sarcopenia (Navarro et al., 2001). Collectively, there is rising recognition that in vivo senescence impacts human physiology and pathology.

1.1.3. Senescence Markers

As cells approach senescence, they undergo specific morphological and molecular changes. Morphologically, senescent cells become flat and enlarged, and accumulate stress granules and vacuoles (Campisi, 1997; Goldstein, 1990). Senescent cells normally display increased activity of the acidic senescence-associated β-galactosidase (SA-β-gal), which is widely used as a marker to identify senescent cells in tissues and cell cultures (Dimri et al., 1995). Senescent cells display significant changes in their nuclear morphology including large nuclei, irregular nuclear envelope, and condensed and disordered chromosomes (Freund et al., 2012; Mehta et al., 2007; Zhang et al., 2007).

The senescence phenotype can also be confirmed by other methods. High levels of ROS are widely used as markers of senescence; as ROS can induce DNA damage. DDR proteins including TP53, p21, γ-H2AX, p16, as well as senescence-associated heterochromatin foci (SAHF) are also major markers of senescence (Narita et al., 2003; Zhang et al., 2005). The localization of HMGB1 can indicate senescence as it enters the extracellular space to stimulate cytokine production through the Toll-like receptor (TLR) 4 pathway (Davalos et al., 2013). Senescent cells can also be identified by the SASP, a phenotype that includes heightened secretion of major chemokines (e.g., IL8, GROA/B/G, MCP2, and HCC4), cytokines (IL6, IL1A, IL1B, and IL7), MMPs (including MMP1/3/10, TIMP1/2), and additional factors (e.g., VEGF, EGF, IGFBP2, NGF, and FGF7) that influence both the immediate cellular microenvironment and distant tissues and organs (Campisi, 2005; Coppe et al., 2006; Kuilman and Peeper, 2009). Together, senescent cells display many specific changes that can serve as markers, although it is important to mention that no single marker is conclusive and thus assessment of multiple markers is needed to identify senescent cells.

1.2. MicroRNAs

MicroRNAs (MiRNAs) are small noncoding RNA molecules (~22 nucleotides long) that generally influence gene expression by binding to subsets of mRNAs with which they share partial complementarity, and suppressing the stability and/or translation of such mRNAs. They are transcribed as long hairpin structures, primary (Pri)-miRNAs, which are processed in the nucleus by Drosha/DGCR8 into precursor (Pre)-miRNAs that are exported to the cytoplasm and further processed by DICER1 into mature single-stranded MiRNAs. Mature MiRNAs are key components of the RNA-inducible silencing complex (RISC) that direct RISC to target mRNAs (Bartel, 2009; Fabian et al., 2010; Kim et al., 2009; Newman and Hammond, 2010). Through their interaction with a wide spectra of mRNAs encoding proteins with diverse functions, MiRNAs have emerged as potent posttranscriptional regulators of several cellular processes, including cell survival, death, division, differentiation, and senescence (Abdelmohsen and Gorospe, 2015; Bushati and Cohen, 2007; Catalanotto et al., 2016; Kloosterman and Plasterk, 2006; Lingor, 2010; Yao, 2016). Like senescence, MiRNAs have been implicated in several disease conditions such as cancer, cardiovascular disease, and neurodegeneration (Saeidimehr et al., 2016; Shah et al., 2016).

The activity of the p53/p21 and p16/RB senescence pathways, as well as the SASP, is governed by the relative levels of the constituents and modulators of these pathways. Among other mechanisms, the levels of the protein components of these signaling systems are strongly regulated by MiRNAs, as discussed in Sections 24.

2. SA-MicroRNAs IN THE P53/P21 PATHWAY

In addition to the finding that suppression of global MiRNA biogenesis induces p53/p21 signaling and senescence, individual MiRNAs have been identified that control the expression levels of protein components of this pathway (Jones and Lal, 2012; Mudhasani et al., 2008). Specifically, we will focus on the SA-MicroRNAs that regulate expression of p53, p21, MDM2, SIRT1, MYC, TERT, and MCD1 (Table 1).

Table 1.

SA-MicroRNAs in the p53/p21 Pathway

MiRNA Target References
MiR-125b, MiR-504, MiR-25, MiR-30d p53 Kumar et al. (2011)
MiR-192, MiR-194, MiR-215, MiR-605 MDM2 Jones and Lal (2012)
MiR-34a, MiR-22, MiR-138, MiR-181a, MiR-217 SIRT1 Jazbutyte et al. (2013), Menghini et al. (2009), Ming et al. (2016), and Yamakuchi (2012)
MiR-34a FoxM1 and MYC Cui et al. (2016) and Xu et al. (2015)
MiR-29c-3p CNOT6 Shang et al. (2016)
MiR-22 CDK6, SIRT1, SP1 Xu et al. (2011)
MiR-195 TERT, SIRT1 Okada et al. (2016)
MiR-22 AKT3 Zheng and Xu (2014)
MiR-200c ZEB1, BMI1 Magenta et al. (2011)
MiR-181a, MiR-181b, MiR-138 SIRT1 Rivetti di Val Cervo et al. (2012) and Zhou et al. (2016)
MiR-130b, MiR-302a, MiR-302b, MiR302c, MiR-302d, MiR-515-3p p21 Borgdorff et al. (2010)
MiR-106b p21 Ivanovska et al. (2008) and Li et al. (2011)
MiR-20a p21 Sokolova et al. (2015)
MiR-17 p21 Gibcus et al. (2011)
MiR-208 p21 Zhang et al. (2011)
MiR-663 p21 Yi et al. (2012)
MiR-519 DUT1, EXO1, RPA2, POLE4 Abdelmohsen et al. (2012) and Marasa et al. (2010)

2.1. p53

MiRNAs MiR-125b, MiR-504, MiR-25, and MiR-30d directly bind p53 mRNA, suppress p53 protein levels, and inhibit senescence (Kumar et al., 2011). For example, lowering MiR-125b levels increased p53 levels and enabled senescence, while elevating MiR-125b levels suppressed p53 abundance and induced apoptosis in human lung fibroblasts responding to stress and developmental cues (Le et al., 2009) (Fig. 1).

Fig. 1.

Fig. 1

SA-MicroRNAs implicated in cellular senescence. Schematic representation of the MicroRNAs that promote or inhibit senescence in three major domains of senescence: the p53/p21 pathway, the p16/RB pathway, and the trait SASP. MicroRNAs that promote senescence are in yellow boxes, MicroRNAs that suppress senescence are in white boxes.

2.2. p21

Several MiRNAs directly suppress p21 expression and thus influence cell senescence and proliferation. For instance, senescence-inhibitory MiRNAs including MiR-130b, MiR-302a, MiR-302b, MiR302c, MiR-302d, and MiR-515-3p attenuate RAS(G12V)-induced senescence of human mammary epithelial cells by preventing the RAS(G12V)-induced upregulation of p21 levels (Borgdorff et al., 2010). MiR-106b reduced p21-mediated cell cycle arrest in prostate cancer cells exhibiting radiation resistance, likely through direct targeting of MiR-106b to the p21 mRNA (Ivanovska et al., 2008; Li et al., 2011). Similarly, MiR-20a was shown to target the 3′UTR of p21 mRNA and blocked p21-mediated growth arrest in TGF-β-responsive colon carcinoma (Sokolova et al., 2015). MiR-208 promoted insulin-induced vascular smooth muscle cell proliferation by down-regulating p21 (Zhang et al., 2011), and MiR-663 suppressed p21 expression and promoted the proliferation and tumorigenesis of nasopharyngeal carcinoma cells (Yi et al., 2012).

Other MiRNAs indirectly enhance p21 expression leading to cell cycle arrest and senescence. For instance, MiR-519a, MiR-519b, and MiR-519c trigger p21-induced senescence by modulating multiple processes including DDR, calcium influx, and autophagy. The impact of MiR-519 on senescence was observed in HeLa cells and WI-38 human diploid fibroblasts (HDFs) (Abdelmohsen et al., 2012; Marasa et al., 2010). The cytoplasmic deadenylase CNOT6 is a component of the deadenylase complex, which contributes to the prevention of cell death and senescence (Mittal et al., 2011). MiR-29c-3p promoted senescence through the p53/p21 and p16/RB pathways by reducing CNOT6 expression levels in human mesenchymal stem cells (Shang et al., 2016).

2.3. MDM2

The murine double minute clone 2 (MDM2) is an oncogene that suppresses the expression of p53 by functioning as a ubiquitin ligase which promotes p53 ubiquitination and degradation by the 26S proteasome (Momand et al., 2000; Piette et al., 1997). Accordingly, the repression of MDM2 expression by MiR-192, MiR-194, MiR-215, and MiR-605 enhanced p53 levels and senescence (Jones and Lal, 2012).

2.4. SIRT1

The silent mating type information regulation 2 homologue 1 (SIRT1) is a deacetylase that regulates metabolic activity in response to cellular stress. SIRT1-mediated deacetylation of p53 suppressed its transcriptional induction of p21 expression (Brooks and Gu, 2009). MiR-34a, MiR-22, MiR-138, MiR-181a, and MiR-217 decrease SIRT1 expression, thereby inducing p53 levels and cell senescence (Jazbutyte et al., 2013; Menghini et al., 2009; Ming et al., 2016; Xu et al., 2011; Yamakuchi, 2012). Interestingly, p53 amplified this effect by transcriptionally inducing the expression of MiR-34a, MiR-34b, and mir-34c, which triggered senescence in various cell systems (Bai et al., 2011; Cui et al., 2016; Kumamoto et al., 2008; Tazawa et al., 2007; Xu et al., 2015).

2.5. MYC

The oncoprotein MYC influences DNA replication, cell cycle progression, and apoptosis. MYC suppresses p21 expression levels, thereby modulating genotoxic and apoptotic responses (Seoane et al., 2002). Several MiRNAs, including MiR-34a, MiR-429, MiR-33b, MiR-126, MiR-136, MiR-145, MiR-449c, and let-7 were found to target the MYC mRNA, although it is not known if they all induce senescence (Deng and Sui, 2013). MiR-34a further impacts senescence via MYC-modulated telomerase activity in human hepatocellular carcinoma (Cui et al., 2016; Xu et al., 2015). Senescence of WI-38 and BJ-1 HDFs triggered by metformin treatment led to increased MiR-429 levels (Cufi et al., 2012), although it is not yet known if MiR-429 targets MYC mRNA in this paradigm.

2.6. TERT

The reverse transcriptase enzyme telomerase (TERT) is essential for maintaining telomere length and thus prevents senescence (Shawi and Autexier, 2008). MiR-195 induced senescence in old mesenchymal stem cells (OMSCs) by directly reducing the levels of TERT and SIRT1. Conversely, inhibiting MiR-195 enhanced the expression of TERT and SIRT1, reduced p53 levels, and thereby delayed senescence (Okada et al., 2016). MiR-1207-5p and MiR-1266 suppressed cell growth and gastric cancer cell invasion by targeting TERT, but it remains to be tested if they induce cell senescence (Chen et al., 2014).

2.7. MDC1

The mediator of DNA damage checkpoint 1 (MDC1) regulates p53 activity and in turn the DDR, DNA repair, cell survival, and cell death (Nakanishi et al., 2007). Accordingly, a rise in MiR-22, which directly represses MDC1 production, hindered the DDR machinery, and induced premature senescence (Coster and Goldberg, 2010; Lee et al., 2015).

3. SA-MicroRNAs ASSOCIATED WITH THE P16/RB PATHWAY

Several MiRNAs have been implicated in p16-mediated senescence. While some MiRNAs influence p16 expression directly through binding to the mRNA, other MiRNAs act upstream to influence p16 levels. Here, we will focus on SA-MiRNAs that affect key components of this pathway, including p16, PRC2, MKK4, and RB (Table 2).

Table 2.

SA-MicroRNAs in the p16/RB Pathway

MicroRNA Target References
MiR-24s p16 Lal et al. (2008) and Philipot et al. (2014)
MiR-26b EED; EZH2 Overhoff et al. (2014)
MiR-181a CBX7 Overhoff et al. (2014)
MiR-210 EED, EZH2, suz12 Overhoff et al. (2014)
MiR-424 EED, suz12 Overhoff et al. (2014)
MiR-138 EZH2 Liang et al. (2013)
MiR15b, MiR-24, MiR-25, MiR-141 MKK4 Marasa et al. (2009)
let-7 p66SHC Xu et al. (2014)
MiR-141 BMI1 Dimri et al. (2013)
MiR-9 CBX7 O’Loghlen et al. (2015)
MiR-29, MiR-30 B-Myb Martinez et al. (2011)
MiR-29c-3p CNOT6 Shang et al. (2016)
MiR-203 E2F3A, E2F3B, ZBP89 Noguchi et al. (2012)
MiR-205 E2F1 Dar et al. (2011)
MiR-449a CCND1, E2F1, E2F3 Noonan et al. (2010), Mao et al. (2016), and Ren et al. (2014)

3.1. p16

The SA-MiRNA MiR-24 binds the 3′UTR of p16 mRNA and suppresses its translation in WI-38 HDFs and in HeLa cells (Lal et al., 2008). While MiR-24 was not found to inhibit senescence, its levels are inversely correlated with the levels of p16 in osteoarthritis-associated senescence (Philipot et al., 2014). In addition, MiR-24 promoted cell proliferation by targeting p16 and the CDK inhibitor p27 in various human cancer cell lines (Giglio et al., 2013). Other MiRNAs suppress p16 expression by targeting p16 mRNA include MiR-300, MiR-514, MiR-663, and MiR-141, although their effect on senescence remains to be studied (Al-Khalaf et al., 2013; Kabir et al., 2015).

3.2. PRC2

p16 is encoded by the INK4A-ARF locus, which is transcriptionally suppressed by the Polycomb-repressive complex 2 (PRC2). Polycomb group (PcG) proteins were found to disassociate from INK4A-ARF locus in senescent cells, allowing p16 levels to increase via increased transcription (Bracken et al., 2007). MiRNAs such as MiR-26b, MiR-181a, MiR-210, and MiR-424 enhanced p16 expression by directly suppressing PRC2 factors like the PcG proteins CBX7, embryonic ectoderm development (EED), enhancer of zeste homologue 2 (EZH2), and suppressor of zeste 12 homologue (Suz12) (Overhoff et al., 2014). Additionally, in human renal cell carcinoma, MiR-138 induced p16 and cell senescence by targeting EZH2 (Liang et al., 2013).

3.3. MKK4

The mitogen-activated protein (MAP) kinase kinase 4 (MKK4) phosphorylates and thereby activates other MAP kinases in response to stress. High levels of MKK4 in senescent cells decreased cell proliferation and triggered senescence (Cunningham et al., 2010). MKK4 mRNA is a target of four MiRNAs (MiR-15b, MiR-24, MiR-25, MiR-141) that jointly suppress MKK4 expression. Downregulation of MKK4 by these MiRNAs decreased p16 levels and delayed senescence, while their coordinated inhibition elevated MKK4 and p16 levels and induced senescence in WI-38 HDFs (Marasa et al., 2009).

3.4. RB

Activation of the RB pathway increased the levels of MiR-29 and MiR-30, which associated with the 3′UTR of the MYBL2 (B-Myb) mRNA, repressed MYBL2 expression and triggered senescence (Martinez et al., 2011; Zhou et al., 2016). While DU-145 prostate cancer cells with mutant RB are resistant to MiR-449a-induced senescence, overexpression of wild-type RB induces senescence by restoring the function of MiR-449a (Noonan et al., 2010). Other studies point to additional MiR-449a targets in the RB pathway; for example, in prostate cancer cells, MiR-449a suppressed E2F1 and induced senescence following ionizing radiation (Mao et al., 2016), and in human lung cancer cells, MiR-449a lowered E2F3 abundance and thus induced senescence (Ren et al., 2014). Ectopic expression of MiR-203 in melanoma cells induced hypophosphorylated RB and elevated other senescence markers by suppressing the expression of E2F3a, E2F3b, and ZBP-89 (Noguchi et al., 2012).

3.5. Other SA-MiRNAs Regulating the 16/RB Pathway

In HDFs, let-7 delayed senescence by reducing the translation of the adaptor protein p66SHC, a key regulator of lifespan in mammals (Xu et al., 2014). In addition, MiR-141 induced p16 levels and consequently senescence by suppressing the production of BMI1 in HDFs (Dimri et al., 2013). In IMR-90 and WI-38 HDFs, MiR-9 was found to promote p16 expression by lowering the abundance of CBX7 (O’Loghlen et al., 2015). Finally, overexpression of MiR-205 in melanoma cells suppressed proliferation and induced markers of senescence including p16 (Dar et al., 2011).

4. SA-MicroRNAs IN SASP

The senescence-associated production and secretion of cytokines, growth factors, and proteases constitutes the SASP (Coppe et al., 2008). This phenotype is governed by transcription factors, RBPs and ncRNAs, including both MiRNAs and lncRNAs. In this section, we focus on MiRNAs that regulate SASP, in particular interleukins (ILs), MMPs, and tumor necrosis factor alpha (TNF) (Table 3).

Table 3.

SA-MicroRNAs in SASP

MicroRNA Target References
MiR-146a/b IRAK1, TLR8 Bhaumik et al. (2009) and Gysler et al. (2016)
MiR-143 MMP13 Osaki et al. (2011)
MiR-335 PTEN Kabir et al. (2016)
MiR-183 ITGB1 Li et al. (2010)
MiR-9 IL6 Zhang et al. (2016a)
MiR-222 MMP1 Liu et al. (2009)
MiR-125b TNF, MMP13 Tili et al. (2007) and Xu et al. (2012)
MiR-152 MMP3 Zheng et al. (2013)
MiR-187 TNF, IL6 Rossato et al. (2012)

4.1. Interleukins

The production and secretion of ILs increases in senescent cells and creates a proinflammatory environment that promotes tumor progression (Coppe et al., 2010). In primary human fibroblasts, MiR-146a and MiR-146b directly suppressed the expression of interleukin 1 receptor-associated kinase 1 (IRAK1), a key regulator of the IL1A (IL-1α) receptor signaling, leading to inhibition of IL6 and IL8 production. Interestingly, high levels of secreted cytokines upregulated MiR-146a/b expression, which in turn prevented excessive SASP activity (Bhaumik et al., 2009). The tumor-suppressor MiR-9 inhibits IL6 expression in cervical adenocarcinoma (Zhang et al., 2016c), and MiR-187, induced by IL10, similarly suppressed the production of IL6 and TNF (Rossato et al., 2012). The circulating MiRNA MiR-21 is a ligand for TLR8 in immune cells, and via this alternative mechanism, it stimulates IL6 and TNF production (Fabbri et al., 2012). MiR-199a suppresses the production of IκB kinase-β (IKBKB), which is required for NF-κB activation, thereby lowering IL6 and IL8 production (Chen et al., 2008).

4.2. Matrix Metalloproteinases

The trait SASP includes the production of numerous MMPs, responsible for degrading the extracellular matrix and cleaving cell surface receptors and ligands. As senescent cells accumulate, the rising MMP activity promotes tissue damage through chronic inflammation and extracellular matrix remodeling caused by SASP (Verma and Hansch, 2007). MMP levels are regulated by different MiRNAs. MiR-222 directly suppresses MMP1 expression in oral tongue squamous cell carcinoma, and further suppresses the production of superoxide dismutase 2 (SOD2), which would otherwise induce MMP1. In addition, MiR-222 indirectly lowers MMP13 levels by downregulating the histone deacetylase HDAC4 (Liu et al., 2009; Song et al., 2015). MiR-152 targets the 3′UTR of MMP3 mRNA, lowering MMP3 expression, and significantly reduces the invasiveness of glioma cells (Zheng et al., 2013). MiR-143 targets the MMP13 3′UTR and suppresses the expression of MMP13. In lung osteosarcoma, the robust down-regulation of MiR-143 leads to a strong increase of MMP13 levels and cancer cell invasiveness (Osaki et al., 2011). The lower levels of MiR-125b in cutaneous squamous cell carcinoma (cSCC) compared with healthy skin allows higher production of MMP13 and accordingly promotes invasion of cSCC (Xu et al., 2012). Finally, MiR-181a-5p directly suppresses MMP14 expression in breast cancer cells (Roth and Cao, 2015).

4.3. TNF

The proinflammatory cytokine TNF is also regulated by MiRNAs. As mentioned earlier, the circulating MiR-21 was capable of binding and activating TLR8 in immune cells, thereby stimulating TNF and IL6 secretion (Fabbri et al., 2012). By contrast, in primary human monocytes, IL10 induced MiR-187, which in turn suppressed TNF and IL6 (Rossato et al., 2012), while in RAW 264.7 macrophages, LPS suppressed MiR-125b, which targets TNF mRNA, and thus enabled TNF production (Tili et al., 2007).

5. SA-MicroRNAs IN PATHOLOGY

As discussed earlier, there is a complex functional connection between senescence and several age-related pathologies. In this section, we discuss some of the major SA-MiRNAs implicated in disease processes.

5.1. Impact on Pathology by SA-MicroRNAs in the p53/p21 Pathway

Many MiRNAs having an impact on the p53/p21 pathway are implicated in diverse pathologies. For example, MiR-181a and MiR-30c synergistically regulate the p53/p21 pathway in diabetes-induced cardiac hypertrophy (Raut et al., 2016). The MiR-34a-mediated reduction of SIRT1, a suppressor of p53 activity, was implicated in tumorigenesis of the colon, lung, and other tissues (Kim et al., 2016; Lai et al., 2012; Misso et al., 2014; Rupaimoole and Slack, 2016; Wang et al., 2013). The Kruppel-like zinc finger protein ZNF224 transcriptionally increased MiR-663a, an MiRNA that lowers p53 and p21 production, and thus ZNF224 induced cell growth and resistance to apoptosis in human breast ductal carcinoma cells (Cho et al., 2016). MiR-621 suppressed production of the p53 inhibitor FBXO11, elevating p53 activity and sensitizing breast cancer cells to apoptosis (Xue et al., 2016), while MiR-24 increased metastasis and invasion of hepatocellular carcinoma cells by lowering p53 levels (Chen et al., 2016b). Collectively, these findings suggest that the network of MiRNAs affecting p53 function can influence pathologies, particularly cancer (Krell et al., 2013).

Recent studies suggest that the MiRNAs affecting p21 expression are particularly tightly involved in cancer. For instance, MiR-95-3p promoted tumorigenesis by targeting and reducing p21 levels in hepatocellular carcinoma (Ye et al., 2016), while inhibition of p21 by MiR-146b enhanced the proliferation of anaplastic thyroid cancer cells (Zhang et al., 2016a). The radiation-induced MiR-208a increased human lung cancer cell proliferation by lowering p21, and both MiR-106b and 208 enhanced radioresistance by lowering p21 expression (Tang et al., 2016; Zheng et al., 2015). In nasopharyngeal carcinoma, MiR-17-5p and MiR-663 promoted cell proliferation and tumorigenesis by lowering p21 abundance (Chen et al., 2016a; Yi et al., 2012), while in esophageal carcinoma, MiR-31 suppressed tumorigenesis in a p21-dependent manner (Ning et al., 2014). The MiRNA cluster MiR-17-92 regulated stem cell potential in MLL leukemia by reducing p21 expression (Wong et al., 2010), while upregulation of p21 by MiR-6734 induced cell cycle arrest and apoptosis in colon cancer cells (Kang et al., 2016). The impact of some of these MiRNAs on senescence remains to be investigated.

5.2. Impact on Pathology by SA-MicroRNAs in the p16/RB Pathway

MiR-24 inhibits p16 expression (Lal et al., 2008) and promoted cell growth, proliferation, and metastasis in gastric cancer cells, nonsmall cell lung cancer, and breast cancer, although a role for p16 in these processes await confirmation (Xu et al., 2013; Zhang et al., 2016b; Zhao et al., 2015). In another disease model, osteoarthritis, MiR-24 was identified as a negative regulator of p16 (Philipot et al., 2014). A rise in p16 levels following the loss of MiR-877-3p suppressed proliferation and tumorigenicity of bladder cancer by arresting cells in the G1 phase (Li et al., 2016b).

5.3. Impact on Pathology by SASP-Associated SA-MicroRNAs

Strong connections between SASP and human pathology are rapidly emerging. Among some prominent examples, lowering MiR-142-3p in macrophages contributed to increasing IL6 levels in aged mice and was proposed to enhance inflammation (Liu et al., 2016), and MiR-218-mediated suppression of the leucine-rich repeat-containing G protein-coupled receptor 4 (LGR4) was proposed to prevent the IL6-induced proliferation and invasion of prostate cancer cells (Li et al., 2016a). In addition, MiR-33a suppressed IL6-induced tumor metastasis of gallbladder cancer by inhibiting the expression of Twist, a transcription factor involved in epithelial-to-mesenchymal transition (Margetts, 2012; Zhang et al., 2016d). Interestingly, p16 suppressed tumorigenesis in breast stromal fibroblasts by repressing the expression and secretion of IL6. This effect was mediated by MiR-146b-5p, which bound to the 3’UTR of IL6 mRNA and suppressed IL6 production (Al-Ansari and Aboussekhra, 2015). In patients with myasthenia gravis, low levels of MiR-181c were found to correlate with high serum levels of IL7, encoded by the MiR-181-target IL7 mRNA (Zhang et al., 2016e). MiR-100-3p and MiR-877-3p suppressed production of IL8 and IL1B, respectively, in mesangial cells from IgA nephropathy patients (Liang et al., 2016). In a number of cultured neuroblastoma cells (SK-N-AS) and glioma cells (U251 and T98G), as well as in patients with glioma, MiR-93-5p suppressed the expression of IL8 and VEGF (Fabbri et al., 2015, 2016). The MiRNA that suppresses IRAK1 expression, MiR-146a, inhibited cell growth, lowered cell migration, and induced apoptosis in nonsmall cell lung cancer (Chen et al., 2013), although the role of IRAK1 in this phenotype was not examined. These reports highlight the extensive functional crosstalk among senescence pathways, inflammation, and disease by MiRNAs that modulate SASP.

The network of MiRNAs affecting senescence-associated MMPs has also been connected to cancer. MiR-143 and MiR-125b suppressed MMP13 expression and inhibited the growth of breast, ovarian, gastric, bladder, and other cancers (Han et al., 2013; Luo et al., 2015; Wang et al., 2016; Wu et al., 2015; Zhai et al., 2016), although the contribution of MMP13 to these effects remains to be confirmed. In human osteosarcoma, MiR-143 lowered the expression of plasminogen activator inhibitor-1 (PAI-1, SERPINE1); the finding that PAI-1 promoted the production and secretion of MMP13, and thereby enhanced invasion and metastasis, is in keeping with the tumor-suppressor function of MiR-143 (Hirahata et al., 2016). Furthermore, as MiR-143 is a circulating MiRNA, it may serve as a biomarker for bladder cancer diagnosis (Motawi et al., 2016). Other MiRNAs like MiR-15b and MiR-152 reduced glioma cell invasion and angiogenesis by suppressing MMP3 and NRP2 (Zheng et al., 2013). MiR-181a-5p suppressed metastases by lowering MMP14 levels in breast cancer cells (Roth and Cao, 2015). It is worth noting that MiR-181a, by lowering SIRT1 and PRC2, was also implicated in activating p53 and inducing p16 levels, respectively, as explained earlier, suggesting that these pathways are coordinately regulated by MiRNAs.

6. CONCLUDING REMARKS AND PERSPECTIVES

The impact of senescent cells on tissue homeostasis is complex and evolves over time, as senescent cells accumulate in the human body with age. Senescent cells are beneficial in some instances, such as during wound healing, muscle regeneration, suppression of pancreatic and liver fibrosis, and prevention of tumorigenesis in young organisms. However, senescent cells can be detrimental in other instances, particularly in the elderly, as they contribute to the development of age-related declines and diseases such as cancer, diabetes, cardiovascular disorders, diminished immunity, and neurodegeneration. Thus, there is rising interest in understanding the molecular regulators of senescence so that we can develop therapeutic approaches that promote its beneficial actions and ameliorate its harmful effects. To help achieve this larger objective, we have reviewed the MiRNAs that promote or suppress senescence.

We discussed SA-MiRNAs and their impact on cellular senescence by focusing on MiRNAs that enhance or diminish the SASP as well as the main senescence pathways, p53/p21 and p16/RB (Tables 13). Some SA-MiRNAs function as direct regulators, for instance, MiR-25 and MiR-30d suppressed senescence by directly targeting the p53 mRNA, while others function as indirect regulators, as shown for MiR-34a and MiR-22, which induced senescence by downregulating SIRT1 levels, which in turn enhanced p53 function, or for MiR-138, which targeted EZH2 leading to transcriptional upregulation of p16 expression. Other SA-MiRNAs, such as MiR-146a and MiR-146b, can modulate SASP, although there is increasing evidence that SA-MiRNAs can coordinately influence multiple senescent traits (Fig. 1).

Importantly, the majority of SA-MiRNAs have been investigated in cell culture models, and so it is important to assess if they change with age and whether they have an impact upon aging and age-related diseases. The development of mouse models of senescence and MiRNA function is starting to enable the systematic analysis of SA-MiRNA function in vivo, although much work is still needed. Understanding the impact of SA-MiRNAs on aging- and age-related diseases will also illuminate their diagnostic and therapeutic potential. For example, SA-MiRNAs in bodily fluids such as blood, urine, and saliva could serve as diagnostic or prognostic markers, perhaps helping to identify disease processes influenced by senescence.

Targeting MiRNAs by gain- or loss-of-function for therapeutic purposes is also an attractive proposition. Despite progress in this area, important challenges remain with the delivery of therapeutic MiRNA-relevant reagents with high specificity and efficacy (Chen et al., 2015). The gain-of-function approach using MiRNA mimics is widely used; however, the outcome should be cautiously interpreted as they may lead to nonspecific changes in gene expression (Jin et al., 2015). The loss-of-function approach using antisense oligos has often been used to inhibit MiRNA activity, but they can also target other sequences with few mismatches resulting in nonspecific effects. To overcome the above limitations, gene knockout would be an ideal choice for MiRNA deletion for functional analysis. Recently, genome-editing technologies based on zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and CRISPR-Cas9-RNA-guided endonucleases (RGENs) have been widely used to knockout genomic loci of interest (Cai and Yang, 2014). Expansion of these technologies to study MiRNAs is emerging (Chang et al., 2016; Kim et al., 2013). Thus, using such new techniques to target MiRNAs will provide an excellent opportunity to further elucidate the role of SA-MiRNAs in vitro and in vivo.

Understanding the full spectrum of transcriptional and posttranscriptional regulators of SA-MiRNAs will further help with accurate targeting of the senescent phenotype. The studies discussed here highlight the efforts underway to exploit senescence regulatory factors toward favorable clinical outcomes in cancer and other diseases influenced by senescent cells.

ACKNOWLEDGMENTS

This work was supported in full by the National Institute on Aging—Intramural Research Program of the National Institutes of Health.

REFERENCES

  1. Abdelmohsen K, Gorospe M, 2015. Noncoding RNA control of cellular senescence. Wiley Interdiscip. Rev. RNA 6, 615–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Abdelmohsen K, Srikantan S, Tominaga K, Kang MJ, Yaniv Y, Martindale JL, Yang XL, Park SS, Becker KG, Subramanian M, Maudsley S, Lal A, Gorospe M, 2012. Growth inhibition by miR-519 via multiple p21-inducing pathways. Mol. Cell. Biol 32, 2530–2548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Acosta JC, O’Loghlen A, Banito A, Guijarro MV, Augert A, Raguz S, Fumagalli M, Da Costa M, Brown C, Popov N, Takatsu Y, Melamed J, di Fagagna FD, Bernard D, Hernando E, Gil J, 2008. Chemokine signaling via the CXCR2 receptor reinforces senescence. Cell 133, 1006–1018. [DOI] [PubMed] [Google Scholar]
  4. Al-Ansari MM, Aboussekhra A, 2015. miR-146b-5p mediates p16-dependent repression of IL-6 and suppresses paracrine procarcinogenic effects of breast stromal fibroblasts. Oncotarget 6, 30006–30016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Al-Khalaf HH, Mohideen P, Nallar SC, Kalvakolanu DV, Aboussekhra A, 2013. The cyclin-dependent kinase inhibitor p16INK4a physically interacts with transcription factor Sp1 and cyclin-dependent kinase 4 to transactivate microRNA-141 and microRNA-146b-5p spontaneously and in response to ultraviolet light-induced DNA damage. J. Biol. Chem 288, 35511–35525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bai XY, Ma YX, Ding R, Fu B, Shi SZ, Chen XM, 2011. miR-335 and miR-34a promote renal senescence by suppressing mitochondrial antioxidative enzymes. J. Am. Soc. Nephrol 22, 1252–1261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Baker DJ, Wijshake T, Tchkonia T, LeBrasseur NK, Childs BG, van de Sluis B, Kirkland JL, van Deursen JM, 2011. Clearance of p16(Ink4a)-positive senescent cells delays ageing-associated disorders. Nature 479, 232–236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bartel DP, 2009. MicroRNAs: target recognition and regulatory functions. Cell 136, 215–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Ben-Porath I, Weinberg RA, 2005. The signals and pathways activating cellular senescence. Int. J. Biochem. Cell Biol 37, 961–976. [DOI] [PubMed] [Google Scholar]
  10. Bhaumik D, Scott GK, Schokrpur S, Patil CK, Orjalo AV, Rodier F, Lithgow GJ, Campisi J, 2009. MicroRNAs miR-146a/b negatively modulate the senescence-associated inflammatory mediators IL-6 and IL-8. Aging 1, 402–411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Borgdorff V, Lleonart ME, Bishop CL, Fessart D, Bergin AH, Overhoff MG, Beach DH, 2010. Multiple microRNAs rescue from Ras-induced senescence by inhibiting p21(Waf1/Cip1). Oncogene 29, 2262–2271. [DOI] [PubMed] [Google Scholar]
  12. Bracken AP, Kleine-Kohlbrecher D, Dietrich N, Pasini D, Gargiulo G, Beekman C, Theilgaard-Monch K, Minucci S, Porse BT, Marine JC, Hansen KH, Helin K, 2007. The polycomb group proteins bind throughout the INK4A-ARF locus and are disassociated in senescent cells. Genes Dev 21, 525–530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Braig M, Lee S, Loddenkemper C, Rudolph C, Peters AHFM, Schlegelberger B, Stein H, Dorken B, Jenuwein T, Schmitt CA, 2005. Oncogene-induced senescence as an initial barrier in lymphoma development. Nature 436, 660–665. [DOI] [PubMed] [Google Scholar]
  14. Brooks CL, Gu W, 2009. How does SIRT1 affect metabolism, senescence and cancer? Nat. Rev. Cancer 9, 123–128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bushati N, Cohen SM, 2007. microRNA functions. Annu. Rev. Cell Dev. Biol 23, 175–205. [DOI] [PubMed] [Google Scholar]
  16. Cai M, Yang Y, 2014. Targeted genome editing tools for disease modeling and gene therapy. Curr. Gene Ther 14, 2–9. [DOI] [PubMed] [Google Scholar]
  17. Campisi J, 1997. Aging and cancer: the double-edged sword of replicative senescence. J. Am. Geriatr. Soc 45, 482–488. [DOI] [PubMed] [Google Scholar]
  18. Campisi J, 2005. Senescent cells, tumor suppression, and organismal aging: good citizens, bad neighbors. Cell 120, 513–522. [DOI] [PubMed] [Google Scholar]
  19. Catalanotto C, Cogoni C, Zardo G, 2016. MicroRNA in control of gene expression: an overview of nuclear functions. Int. J. Mol. Sci 17, pii: E1712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Chang H, Yi B, Ma RX, Zhang XG, Zhao HY, Xi YG, 2016. CRISPR/cas9, a novel genomic tool to knock down microRNA in vitro and in vivo. Sci. Rep 6, 22312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Chen R, Alvero AB, Silasi DA, Kelly MG, Fest S, Visintin I, Leiser A, Schwartz PE, Rutherford T, Mor G, 2008. Regulation of IKKbeta by miR-199a affects NF-kappaB activity in ovarian cancer cells. Oncogene 27, 4712–4723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Chen G, Umelo IA, Lv SS, Teugels E, Fostier K, Kronenberger P, Dewaele A, Sadones J, Geers C, De Greve J, 2013. miR-146a inhibits cell growth, cell migration and induces apoptosis in non-small cell lung cancer cells. Plos One 8, e60317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chen L, Lu MH, Zhang D, Hao NB, Fan YH, Wu YY, Wang SM, Xie R, Fang DC, Zhang H, Hu CJ, Yang SM, 2014. miR-1207–5p and miR-1266 suppress gastric cancer growth and invasion by targeting telomerase reverse transcriptase. Cell Death Dis 5, e1034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chen YC, Gao DY, Huang L, 2015. In vivo delivery of miRNAs for cancer therapy: challenges and strategies. Adv. Drug Deliver. Rev 81, 128–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Chen C, Lu Z, Yang J, Hao W, Qin Y, Wang H, Xie C, Xie R, 2016a. MiR-17-5p promotes cancer cell proliferation and tumorigenesis in nasopharyngeal carcinoma by targeting p21. Cancer Med 5, 3489–3499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chen L, Luo L, Chen W, Xu HX, Chen F, Chen LZ, Zeng WT, Chen JS, Huang XH, 2016b. MicroRNA-24 increases hepatocellular carcinoma cell metastasis and invasion by targeting p53: miR-24 targeted p53. Biomed. Pharmacother 84, 1113–1118. [DOI] [PubMed] [Google Scholar]
  27. Childs BG, Durik M, Baker DJ, van Deursen JM, 2015. Cellular senescence in aging and age-related disease: from mechanisms to therapy. Nat. Med 21, 1424–1435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chinta SJ, Woods G, Rane A, Demaria M, Campisi J, Andersen JK, 2015. Cellular senescence and the aging brain. Exp. Gerontol 68, 3–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Cho JG, Park S, Lim CH, Kim HS, Song SY, Roh TY, Sung JH, Suh W, Ham SJ, Lim KH, Park SG, 2016. ZNF224, Kruppel like zinc finger protein, induces cell growth and apoptosis-resistance by down-regulation of p21 and p53 via miR-663a. Oncotarget 7, 31177–31190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Choi J, Shendrik I, Peacocke M, Peehl D, Buttyan R, Ikeguchi EF, Katz AE, Benson MC, 2000. Expression of senescence-associate beta-galactosidase in enlarged prostates from men with benign prostatic hyperplasia. Urology 56, 160–166. [DOI] [PubMed] [Google Scholar]
  31. Collado M, Gil J, Efeyan A, Guerra C, Schuhmacher AJ, Barradas M, Benguria A, Zaballos A, Flores JM, Barbacid M, Beach D, Serrano M, 2005. Tumour biology—senescence in premalignant tumours. Nature 436, 642. [DOI] [PubMed] [Google Scholar]
  32. Coppe JP, Kauser K, Campisi J, Beausejour CM, 2006. Secretion of vascular endothelial growth factor by primary human fibroblasts at senescence. J. Biol. Chem 281, 29568–29574. [DOI] [PubMed] [Google Scholar]
  33. Coppe JP, Patil CK, Rodier F, Sun Y, Munoz DP, Goldstein J, Nelson PS, Desprez PY, Campisi J, 2008. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol 6, 2853–2868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Coppe JP, Desprez PY, Krtolica A, Campisi J, 2010. The senescence-associated secretory phenotype: the dark side of tumor suppression. Annu. Rev. Pathol. Mech. Dis 5, 99–118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Coster G, Goldberg M, 2010. The cellular response to DNA damage: a focus on MDC1 and its interacting proteins. Nucleus 1, 166–178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Cristofalo VJ, Lorenzini A, Allen RG, Torres C, Tresini M, 2004. Replicative senescence: a critical review. Mech. Ageing Dev 125, 827–848. [DOI] [PubMed] [Google Scholar]
  37. Cufi S, Vazquez-Martin A, Oliveras-Ferraros C, Quirantes R, Segura-Carretero A, Micol V, Joven J, Bosch-Barrera J, Del Barco S, Martin-Castillo B, Vellon L, Menendez JA, 2012. Metformin lowers the threshold for stress-induced senescence: a role for the microRNA-200 family and miR-205. Cell Cycle 11, 1235–1246. [DOI] [PubMed] [Google Scholar]
  38. Cui H, Ge J, Xie N, Banerjee S, Zhou Y, Antony VB, Thannickal VJ, Liu G, 2016. miR-34a inhibits lung fibrosis by inducing lung fibroblast senescence. Am. J. Respir. Cell Mol. Biol 56, 168–178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Cunningham SC, Gallmeier E, Kern SE, 2010. MKK4 as oncogene or tumor supressor: in cancer and senescence, the story’s getting old. Aging 2, 752–753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Dankort D, Curley DP, Cartlidge RA, Nelson B, Karnezis AN, Damsky WE, You MJ, DePinho RA, McMahon M, Bosenberg M, 2009. Braf(V600E) cooperates with Pten loss to induce metastatic melanoma. Nat. Genet 41, 544–552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Dar AA, Majid S, de Semir D, Nosrati M, Bezrookove V, Kashani-Sabet M, 2011. miRNA-205 suppresses melanoma cell proliferation and induces senescence via regulation of E2F1 protein. J. Biol. Chem 286, 16606–16614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Davalos AR, Kawahara M, Malhotra GK, Schaum N, Huang JH, Ved U, Beausejour CM, Coppe JP, Rodier F, Campisi J, 2013. p53-dependent release of Alarmin HMGB1 is a central mediator of senescent phenotypes. J. Cell Biol 201, 613–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Demaria M, Ohtani N, Youssef SA, Rodier F, Toussaint W, Mitchell JR, Laberge RM, Vijg J, Van Steeg H, Dolle MET, Hoeijmakers JHJ, de Bruin A, Hara E, Campisi J, 2014. An essential role for senescent cells in optimal wound healing through secretion of PDGF-AA. Dev. Cell 31, 722–733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Deng G, Sui G, 2013. Noncoding RNA in oncogenesis: a new era of identifying key players. Int. J. Mol. Sci 14, 18319–18349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Dimri GP, Lee XH, Basile G, Acosta M, Scott C, Roskelley C, Medrano EE, Linskens M, Rubelj I, Pereirasmith O, Peacocke M, Campisi J, 1995. A biomarker that identifies senescent human-cells in culture and in aging skin in-vivo. Proc. Natl. Acad. Sci. U.S.A 92, 9363–9367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Dimri M, Carroll JD, Cho JH, Dimri GP, 2013. microRNA-141 regulates BMI1 expression and induces senescence in human diploid fibroblasts. Cell Cycle 12, 3537–3546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Fabbri M, Paone A, Calore F, Galli R, Gaudio E, Santhanam R, Lovat F, Fadda P, Mao C, Nuovo GJ, Zanesi N, Crawford M, Ozer GH, Wernicke D, Alder H, Caligiuri MA, Nana-Sinkam P, Perrotti D, Croce CM, 2012. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc. Natl. Acad. Sci. U.S.A 109, E2110–E2116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Fabbri E, Brognara E, Montagner G, Ghimenton C, Eccher A, Cantu C, Khalil S, Bezzerri V, Provezza L, Bianchi N, Finotti A, Borgatti M, Moretto G, Chilosi M, Cabrini G, Gambari R, 2015. Regulation of IL-8 gene expression in gliomas by microRNA miR-93. BMC Cancer 15, 661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Fabbri E, Montagner G, Bianchi N, Finotti A, Borgatti M, Lampronti I, Cabrini G, Gambari R, 2016. MicroRNA miR-93–5p regulates expression of IL-8 and VEGF in neuroblastoma SK-N-AS cells. Oncol. Rep 35, 2866–2872. [DOI] [PubMed] [Google Scholar]
  50. Fabian MR, Sonenberg N, Filipowicz W, 2010. Regulation of mRNA translation and stability by microRNAs. Annu. Rev. Biochem 79, 351–379. [DOI] [PubMed] [Google Scholar]
  51. Freund A, Orjalo AV, Desprez PY, Campisi J, 2010. Inflammatory networks during cellular senescence: causes and consequences. Trends Mol. Med 16, 238–246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Freund A, Laberge RM, Demaria M, Campisi J, 2012. Lamin B1 loss is a senescence-associated biomarker. Mol. Biol. Cell 23, 2066–2075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Gey C, Seeger K, 2013. Metabolic changes during cellular senescence investigated by proton NMR-spectroscopy. Mech. Ageing Dev 134, 130–138. [DOI] [PubMed] [Google Scholar]
  54. Giglio S, Cirombella R, Amodeo R, Portaro L, Lavra L, Vecchione A, 2013. MicroRNA miR-24 promotes cell proliferation by targeting the CDKs inhibitors p27Kip1 and p16INK4a. J. Cell. Physiol 228, 2015–2023. [DOI] [PubMed] [Google Scholar]
  55. Goldstein S, 1990. Replicative senescence: the human fibroblast comes of age. Science 249, 1129–1133. [DOI] [PubMed] [Google Scholar]
  56. Gorospe M, Abdelmohsen K, 2011. MicroRegulators come of age in senescence. Trends Genet 27, 233–241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Han Y, Liu Y, Zhang H, Wang T, Diao R, Jiang Z, Gui Y, Cai Z, 2013. Has-miR-125b suppresses bladder cancer development by down-regulating oncogene SIRT7 and oncogenic long non-coding RNA MALAT1. FEBS Lett 587, 3875–3882. [PubMed] [Google Scholar]
  58. Hayflick L, 1965. The limited in vitro lifetime of human diploid cell strains. Exp. Cell Res 37, 614–636. [DOI] [PubMed] [Google Scholar]
  59. Hayflick L, 1992. Aging, longevity, and immortality in vitro. Exp. Gerontol 27, 363–368. [DOI] [PubMed] [Google Scholar]
  60. Hayflick L, Moorhead PS, 1961. The serial cultivation of human diploid cell strains. Exp. Cell Res 25, 585–621. [DOI] [PubMed] [Google Scholar]
  61. Herbig U, Jobling WA, Chen BP, Chen DJ, Sedivy JM, 2004. Telomere shortening triggers senescence of human cells through a pathway involving ATM, p53, and p21 (CIP1), but not p16(INK4a). Mol. Cell 14, 501–513. [DOI] [PubMed] [Google Scholar]
  62. Herbig U, Ferreira M, Condel L, Carey D, Sedivy JM, 2006. Cellular senescence in aging primates. Science 311, 1257. [DOI] [PubMed] [Google Scholar]
  63. Hirahata M, Osaki M, Kanda Y, Sugimoto Y, Yoshioka Y, Kosaka N, Takeshita F, Fujiwara T, Kawai A, Ito H, Ochiya T, Okada F, 2016. PAI-1, a target gene of miR-143, regulates invasion and metastasis by upregulating MMP-13 expression of human osteosarcoma. Cancer Med 5, 892–902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Ivanovska I, Ball AS, Diaz RL, Magnus JF, Kibukawa M, Schelter JM, Kobayashi SV, Lim L, Burchard J, Jackson AL, Linsley PS, Cleary MA, 2008. MicroRNAs in the miR-106b family regulate p21/CDKN1A and promote cell cycle progression. Mol. Cell. Biol 28, 2167–2174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Jazbutyte V, Fiedler J, Kneitz S, Galuppo P, Just A, Holzmann A, Bauersachs J, Thum T, 2013. MicroRNA-22 increases senescence and activates cardiac fibroblasts in the aging heart. Age 35, 747–762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Jeyapalan JC, Ferreira M, Sedivy JA, Herbig U, 2007. Accumulation of senescent cells in mitotic tissue of aging primates. Mech. Ageing Dev 128, 36–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Jin HY, Gonzalez-Martin A, Miletic AV, Lai MY, Knight S, Sabouri-Ghomi M, Head SR, Macauley MS, Rickert RC, Xiao CC, 2015. Transfection of microRNA mimics should be used with caution. Front. Genet 6, 340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Jones M, Lal A, 2012. MicroRNAs, wild-type and mutant p53: more questions than answers. RNA Biol 9, 781–791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Kabir FML, DeInnocentes P, Bird RC, 2015. Altered microRNA expression profiles and regulation of INK4A/CDKN2A tumor suppressor genes in canine breast cancer models. J. Cell. Biochem 116, 2956–2969. [DOI] [PubMed] [Google Scholar]
  70. Kang MR, Park KH, Yang JO, Lee CW, Oh SJ, Yun J, Lee MY, Han SB, Kang JS, 2016. miR-6734 up-regulates p21 gene expression and induces cell cycle arrest and apoptosis in colon cancer cells. PloS One 11, e0160961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Kim VN, Han J, Siomi MC, 2009. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell. Biol 10, 126–139. [DOI] [PubMed] [Google Scholar]
  72. Kim YK, Wee G, Park J, Kim J, Baek D, Kim JS, Kim VN, 2013. TALEN-based knockout library for human microRNAs. Nat. Struct. Mol. Biol 20, 1458–1464. [DOI] [PubMed] [Google Scholar]
  73. Kim YH, Lee WK, Lee EB, Son JW, Kim DS, Park JY, 2016. Combined effect of metastasis-related microRNA, miR-34 and miR-124 family, methylation on prognosis of non-small-cell lung cancer. Clin. Lung Cancer 18, e13–e20. [DOI] [PubMed] [Google Scholar]
  74. Kloosterman WP, Plasterk RHA, 2006. The diverse functions of MicroRNAs in animal development and disease. Dev. Cell 11, 441–450. [DOI] [PubMed] [Google Scholar]
  75. Krell J, Frampton AE, Colombo T, Gall TM, De Giorgio A, Harding V, Stebbing J, Castellano L, 2013. The p53 miRNA interactome and its potential role in the cancer clinic. Epigenomics 5, 417–428. [DOI] [PubMed] [Google Scholar]
  76. Krishnamurthty J, Torrice C, Ramsey MR, Kovalev GI, Al-Regaiey K, Su LS, Sharpless NE, 2004. Ink4a/Arf expression is a biomarker of aging. J. Clin. Invest 114, 1299–1307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Krishnamurthy J, Ramsey MR, Ligon KL, Torrice C, Koh A, Bonner-Weir S, Sharpless NE, 2006. p16INK4a induces an age-dependent decline in islet regenerative potential. Nature 443, 453–457. [DOI] [PubMed] [Google Scholar]
  78. Krizhanovsky V, Yon M, Dickins RA, Hearn S, Simon J, Miething C, Yee H, Zender L, Lowe SW, 2008. Senescence of activated stellate cells limits liver fibrosis. Cell 134, 657–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Kuilman T, Peeper DS, 2009. Senescence-messaging secretome: SMS-ing cellular stress. Nat. Rev. Cancer 9, 81–94. [DOI] [PubMed] [Google Scholar]
  80. Kuilman T, Michaloglou C, Vredeveld LCW, Douma S, van Doom R, Desmet CJ, Aarden LA, Mooi WJ, Peeper DS, 2008. Oncogene-induced senescence relayed by an interleukin-dependent inflammatory network. Cell 133, 1019–1031. [DOI] [PubMed] [Google Scholar]
  81. Kuilman T, Michaloglou C, Mooi WJ, Peeper DS, 2010. The essence of senescence. Genes Dev 24, 2463–2479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Kumamoto K, Spillare EA, Fujita K, Horikawa I, Yamashita T, Appella E, Nagashima M, Takenoshita S, Yokota J, Harris CC, 2008. Nutlin-3a activates p53 to both down-regulate inhibitor of growth 2 and up-regulate mir-34a, mir-34b, and mir-34c expression, and induce senescence. Cancer Res 68, 3193–3203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Kumar M, Lu Z, Takwi AA, Chen W, Callander NS, Ramos KS, Young KH, Li Y, 2011. Negative regulation of the tumor suppressor p53 gene by microRNAs. Oncogene 30, 843–853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Lai X, Wolkenhauer O, Vera J, 2012. Modeling miRNA regulation in cancer signaling systems: miR-34a regulation of the p53/Sirt1 signaling module. Methods Mol. Biol 880, 87–108. [DOI] [PubMed] [Google Scholar]
  85. Lal A, Kim HH, Abdelmohsen K, Kuwano Y, Pullmann R, Srikantan S, Subrahmanyam R, Martindale JL, Yang XL, Ahmed F, Navarro F, Dykxhoorn D, Lieberman J, Gorospe M, 2008. p16(INK4a) translation suppressed by miR-24. Plos One 3, e1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Le MTN, Teh C, Shyh-Chang N, Xie HM, Zhou BY, Korzh V, Lodish HF, Lim B, 2009. MicroRNA-125b is a novel negative regulator of p53. Genes Dev 23, 862–876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Lee JH, Park SJ, Jeong SY, Kim MJ, Jun S, Lee HS, Chang IY, Lim SC, Yoon SP, Yong J, You HJ, 2015. MicroRNA-22 suppresses DNA repair and promotes genomic instability through targeting of MDC1. Cancer Res 75, 1298–1310. [DOI] [PubMed] [Google Scholar]
  88. Li B, Shi XB, Nori D, Chao CK, Chen AM, Valicenti R, White Rde V, 2011. Down-regulation of microRNA 106b is involved in p21-mediated cell cycle arrest in response to radiation in prostate cancer cells. Prostate 71, 567–574. [DOI] [PubMed] [Google Scholar]
  89. Li F, Gu C, Tian F, Jia Z, Meng Z, Ding Y, Yang J, 2016a. MiR-218 impedes IL-6-induced prostate cancer cell proliferation and invasion via suppression of LGR4 expression. Oncol. Rep 35, 2859–2865. [DOI] [PubMed] [Google Scholar]
  90. Li S, Zhu Y, Liang Z, Wang X, Meng S, Xu X, Xu X, Wu J, Ji A, Hu Z, Lin Y, Chen H, Mao Y, Wang W, Zheng X, Liu B, Xie L, 2016b. Up-regulation of p16 by miR-877–3p inhibits proliferation of bladder cancer. Oncotarget 7, 51773–51783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Liang J, Zhang Y, Jiang G, Liu Z, Xiang W, Chen X, Chen Z, Zhao J, 2013. MiR-138 induces renal carcinoma cell senescence by targeting EZH2 and is down-regulated in human clear cell renal cell carcinoma. Oncol. Res 21, 83–91. [DOI] [PubMed] [Google Scholar]
  92. Liang Y, Zhao G, Tang L, Zhang J, Li T, Liu Z, 2016. MiR-100-3p and miR-877-3p regulate overproduction of IL-8 and IL-1beta in mesangial cells activated by secretory IgA from IgA nephropathy patients. Exp. Cell Res 347, 312–321. [DOI] [PubMed] [Google Scholar]
  93. Lingor P, 2010. Regulation of Cell Death and Survival by RNA Interference—The Roles of miRNA and siRNA. In: Apoptosome: An up-and-Coming Therapeutical Tool, Springer, New York, pp. 95–117. [Google Scholar]
  94. Liu X, Yu J, Jiang L, Wang A, Shi F, Ye H, Zhou X, 2009. MicroRNA-222 regulates cell invasion by targeting matrix metalloproteinase 1 (MMP1) and manganese superoxide dismutase 2 (SOD2) in tongue squamous cell carcinoma cell lines. Cancer Genomics Proteomics 6, 131–139. [PMC free article] [PubMed] [Google Scholar]
  95. Liu Y, Song X, Meng S, Jiang M, 2016. Downregulated expression of miR-142–3p in macrophages contributes to increased IL-6 levels in aged mice. Mol. Immunol 80, 11–16. [DOI] [PubMed] [Google Scholar]
  96. Loaiza N, Demaria M, 2016. Cellular senescence and tumor promotion: is aging the key? Biochim. Biophys. Acta 1865, 155–167. [DOI] [PubMed] [Google Scholar]
  97. Luo XG, Ding JQ, Chen SD, 2010. Microglia in the aging brain: relevance to neurodegeneration. Mol. Neurodegener 5, 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Luo S, Wang J, Ma Y, Yao Z, Pan H, 2015. PPARgamma inhibits ovarian cancer cells proliferation through upregulation of miR-125b. Biochem. Biophys. Res. Commun 462, 85–90. [DOI] [PubMed] [Google Scholar]
  99. Mao A, Liu Y, Wang Y, Zhao Q, Zhou X, Sun C, Di C, Si J, Gan L, Zhang H, 2016. miR-449a enhances radiosensitivity through modulating pRb/E2F1 in prostate cancer cells. Tumour Biol 37, 4831–4840. [DOI] [PubMed] [Google Scholar]
  100. Marasa BS, Srikantan S, Masuda K, Abdelmohsen K, Kuwano Y, Yang X, Martindale JL, Rinker-Schaeffer CW, Gorospe M, 2009. Increased MKK4 abundance with replicative senescence is linked to the joint reduction of multiple microRNAs. Sci. Signal 2, ra69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Marasa BS, Srikantan S, Martindale JL, Kim MM, Lee EK, Gorospe M, Abdelmohsen K, 2010. MicroRNA profiling in human diploid fibroblasts uncovers miR-519 role in replicative senescence. Aging 2, 333–343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Margetts PJ, 2012. Twist: a new player in the epithelial-mesenchymal transition of the peritoneal mesothelial cells. Nephrol. Dial. Transplant 27, 3978–3981. [DOI] [PubMed] [Google Scholar]
  103. Martinez I, Cazalla D, Almstead LL, Steitz JA, DiMaio D, 2011. miR-29 and miR-30 regulate B-Myb expression during cellular senescence. Proc. Natl. Acad. Sci. U.S.A 108, 522–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Massague J, 2008. TGF beta in cancer. Cell 134, 215–230.18662538 [Google Scholar]
  105. Mehta IS, Figgitt M, Clements CS, Kill IR, Bridger JM, 2007. Alterations to nuclear architecture and genome behavior in senescent cells. Ann. N. Y. Acad. Sci 1100, 250–263. [DOI] [PubMed] [Google Scholar]
  106. Menghini R, Casagrande V, Cardellini M, Martelli E, Terrinoni A, Amati F, Vasa-Nicotera M, Ippoliti A, Novelli G, Melino G, Lauro R, Federici M, 2009. MicroRNA 217 modulates endothelial cell senescence via silent information regulator 1. Circulation 120, 1524–1532. [DOI] [PubMed] [Google Scholar]
  107. Michaloglou C, Vredeveld LCW, Soengas MS, Denoyelle C, Kuilman T, van der Horst CMAM, Majoor DM, Shay JW, Mooi WJ, Peeper DS, 2005. BRAF (E600)-associated senescence-like cell cycle arrest of human naevi. Nature 436, 720–724. [DOI] [PubMed] [Google Scholar]
  108. Ming GF, Wu K, Hu K, Chen Y, Xiao J, 2016. NAMPT regulates senescence, proliferation, and migration of endothelial progenitor cells through the SIRT1 AS lncRNA/miR-22/SIRT1 pathway. Biochem. Biophys. Res. Commun 478, 1382–1388. [DOI] [PubMed] [Google Scholar]
  109. Misso G, Di Martino MT, De Rosa G, Farooqi AA, Lombardi A, Campani V, Zarone MR, Gulla A, Tagliaferri P, Tassone P, Caraglia M, 2014. Mir-34: a new weapon against cancer? Mol. Ther. Nucleic Acids 3, e194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Mittal S, Aslam A, Doidge R, Medica R, Winkler GS, 2011. The Ccr4a (CNOT6) and Ccr4b (CNOT6L) deadenylase subunits of the human Ccr4-Not complex contribute to the prevention of cell death and senescence. Mol. Biol. Cell 22, 748–758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Molofsky AV, Slutsky SG, Joseph NM, He S, Pardal R, Krishnamurthy J, Sharpless NE, Morrison SJ, 2006. Increasing p16INK4a expression decreases fore-brain progenitors and neurogenesis during ageing. Nature 443, 448–452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Momand J, Wu HH, Dasgupta G, 2000. MDM2—master regulator of the p53 tumor suppressor protein. Gene 242, 15–29. [DOI] [PubMed] [Google Scholar]
  113. Motawi TK, Rizk SM, Ibrahim TM, Ibrahim IA, 2016. Circulating microRNAs, miR-92a, miR-100 and miR-143, as non-invasive biomarkers for bladder cancer diagnosis. Cell Biochem. Funct 34, 142–148. [DOI] [PubMed] [Google Scholar]
  114. Mudhasani R, Zhu Z, Hutvagner G, Eischen CM, Lyle S, Hall LL, Lawrence JB, Imbalzano AN, Jones SN, 2008. Loss of miRNA biogenesis induces p19Arf-p53 signaling and senescence in primary cells. J. Cell Biol 181, 1055–1063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Nakanishi M, Ozaki T, Yamamoto H, Hanamoto T, Kikuchi H, Furuya K, Asaka M, Delia D, Nakagawara A, 2007. NFBD1/MDC1 associates with p53 and regulates its function at the crossroad between cell survival and death in response to DNA damage. J. Biol. Chem 282, 22993–23004. [DOI] [PubMed] [Google Scholar]
  116. Narita M, Nunez S, Heard E, Narita M, Lin AW, Hearn SA, Spector DL, Hannon GJ, Lowe SW, 2003. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell 113, 703–716. [DOI] [PubMed] [Google Scholar]
  117. Navarro A, Lopez-Cepero JM, del Pino MJS, 2001. Skeletal muscle and aging. Front. Biosci 6, D26–D44. [DOI] [PubMed] [Google Scholar]
  118. Newman MA, Hammond SM, 2010. Emerging paradigms of regulated microRNA processing. Gene Dev 24, 1086–1092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Ning Z, Zhu H, Li F, Liu Q, Liu G, Tan T, Zhang B, Chen S, Li G, Huang D, Meltzer SJ, Zhang H, 2014. Tumor suppression by miR-31 in esophageal carcinoma is p21-dependent. Genes Cancer 5, 436–444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Noguchi S, Mori T, Otsuka Y, Yamada N, Yasui Y, Iwasaki J, Kumazaki M, Maruo K, Akao Y, 2012. Anti-oncogenic microRNA-203 induces senescence by targeting E2F3 protein in human melanoma cells. J. Biol. Chem 287, 11769–11777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Noonan EJ, Place RF, Basak S, Pookot D, Li LC, 2010. miR-449a causes Rb-dependent cell cycle arrest and senescence in prostate cancer cells. Oncotarget 1, 349–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Ohtani N, Takahashi A, Mann DJ, Hara E, 2012. Cellular senescence: a double-edged sword in the fight against cancer. Exp. Dermatol 21 (Suppl. 1), 1–4. [DOI] [PubMed] [Google Scholar]
  123. Okada M, Kim HW, Matsu-ura K, Wang YG, Xu M, Ashraf M, 2016. Abrogation of age-induced microRNA-195 rejuvenates the senescent mesenchymal stem cells by reactivating telomerase. Stem Cells 34, 148–159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. O’Loghlen A, Brookes S, Martin N, Rapisarda V, Peters G, Gil J, 2015. CBX7 and miR-9 are part of an autoregulatory loop controlling p16(INK) (4a). Aging Cell 14, 1113–1121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Osaki M, Takeshita F, Sugimoto Y, Kosaka N, Yamamoto Y, Yoshioka Y, Kobayashi E, Yamada T, Kawai A, Inoue T, Ito H, Oshimura M, Ochiya T, 2011. MicroRNA-143 regulates human osteosarcoma metastasis by regulating matrix metalloprotease-13 expression. Mol. Ther 19, 1123–1130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Overhoff MG, Garbe JC, Koh J, Stampfer MR, Beach DH, Bishop CL, 2014. Cellular senescence mediated by p16INK4A-coupled miRNA pathways. Nucleic Acids Res 42, 1606–1618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Philipot D, Guerit D, Platano D, Chuchana P, Olivotto E, Espinoza F, Dorandeu A, Pers YM, Piette J, Borzi RM, Jorgensen C, Noel D, Brondello JM, 2014. p16(INK4a) and its regulator miR-24 link senescence and chondrocyte terminal differentiation-associated matrix remodeling in osteoarthritis. Arthritis Res. Ther 16, R58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Piette J, Neel H, Marechal V, 1997. Mdm2: keeping p53 under control. Oncogene 15, 1001–1010. [DOI] [PubMed] [Google Scholar]
  129. Raut SK, Singh GB, Rastogi B, Saikia UN, Mittal A, Dogra N, Singh S, Prasad R, Khullar M, 2016. miR-30c and miR-181a synergistically modulate p53-p21 pathway in diabetes induced cardiac hypertrophy. Mol. Cell. Biochem 417, 191–203. [DOI] [PubMed] [Google Scholar]
  130. Ren XS, Yin MH, Zhang X, Wang Z, Feng SP, Wang GX, Luo YJ, Liang PZ, Yang XQ, He JX, Zhang BL, 2014. Tumor-suppressive microRNA-449a induces growth arrest and senescence by targeting E2F3 in human lung cancer cells. Cancer Lett 344, 195–203. [DOI] [PubMed] [Google Scholar]
  131. Rossato M, Curtale G, Tamassia N, Castellucci M, Mori L, Gasperini S, Mariotti B, De Luca M, Mirolo M, Cassatella MA, Locati M, Bazzoni F, 2012. IL-10-induced microRNA-187 negatively regulates TNF-alpha, IL-6, and IL-12p40 production in TLR4-stimulated monocytes. Proc. Natl. Acad. Sci. U.S.A 109, E3101–E3110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Rosso A, Balsamo A, Gambino R, Dentelli P, Falcioni R, Cassader M, Pegoraro L, Pagano G, Brizzi MF, 2006. p53 mediates the accelerated onset of senescence of endothelial progenitor cells in diabetes. J. Biol. Chem 281, 4339–4347. [DOI] [PubMed] [Google Scholar]
  133. Roth E, Cao J, 2015. miR-181 suppresses metastasis via MMP-14. Aging 7, 740–741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Rupaimoole R, Slack FJ, 2016. A role for miR-34 in colon cancer stem cell homeostasis. Stem Cell Investig 3, 42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Saeidimehr S, Ebrahimi A, Saki N, Goodarzi P, Rahim F, 2016. MicroRNA-based linkage between aging and cancer: from epigenetics view point. Cell J 18, 117–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Seoane J, Le HV, Massague J, 2002. Myc suppression of the p21(Cip1) Cdk inhibitor influences the outcome of the p53 response to DNA damage. Nature 419, 729–734. [DOI] [PubMed] [Google Scholar]
  137. Shah MY, Ferrajoli A, Sood AK, Lopez-Berestein G, Calin GA, 2016. microRNA therapeutics in cancer—an emerging concept. EBioMedicine 12, 34–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Shang J, Yao Y, Fan X, Shangguan L, Li J, Liu H, Zhou Y, 2016. miR-29c-3p promotes senescence of human mesenchymal stem cells by targeting CNOT6 through p53-p21 and p16-pRB pathways. Biochim. Biophys. Acta 1863, 520–532. [DOI] [PubMed] [Google Scholar]
  139. Shawi M, Autexier C, 2008. Telomerase, senescence and ageing. Mech. Ageing Dev 129, 3–10. [DOI] [PubMed] [Google Scholar]
  140. Sokolova V, Fiorino A, Zoni E, Crippa E, Reid JF, Gariboldi M, Pierotti MA, 2015. The effects of miR-20a on p21: two mechanisms blocking growth arrest in TGF-beta-responsive colon carcinoma. J. Cell. Physiol 230, 3105–3114. [DOI] [PubMed] [Google Scholar]
  141. Song J, Jin EH, Kim D, Kim KY, Chun CH, Jin EJ, 2015. MicroRNA-222 regulates MMP-13 via targeting HDAC-4 during osteoarthritis pathogenesis. BBA Clin 3, 79–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Sousa-Victor P, Gutarra S, Garcia-Prat L, Rodriguez-Ubreva J, Ortet L, Ruiz-Bonilla V, Jardi M, Ballestar E, Gonzalez S, Serrano AL, Perdiguero E, Munoz-Canoves P, 2014. Geriatric muscle stem cells switch reversible quiescence into senescence. Nature 506, 316–321. [DOI] [PubMed] [Google Scholar]
  143. Tang Y, Cui Y, Li Z, Jiao Z, Zhang Y, He Y, Chen G, Zhou Q, Wang W, Zhou X, Luo J, Zhang S, 2016. Radiation-induced miR-208a increases the proliferation and radioresistance by targeting p21 in human lung cancer cells. J. Exp. Clin. Cancer Res 35, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Tazawa H, Tsuchiya N, Izumiya M, Nakagama H, 2007. Tumor-suppressive miR-34a induces senescence-like growth arrest through modulation of the E2F pathway in human colon cancer cells. Proc. Natl. Acad. Sci. U.S.A 104, 15472–15477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Tili E, Michaille JJ, Cimino A, Costinean S, Dumitru CD, Adair B, Fabbri M, Alder H, Liu CG, Calin GA, Croce CM, 2007. Modulation of miR-155 and miR-125b levels following lipopolysaccharide/TNF-alpha stimulation and their possible roles in regulating the response to endotoxin shock. J. Immunol 179, 5082–5089. [DOI] [PubMed] [Google Scholar]
  146. Verma RP, Hansch C, 2007. Matrix metalloproteinases (MMPs): chemical-biological functions and (Q)SARs. Bioorg. Med. Chem 15, 2223–2268. [DOI] [PubMed] [Google Scholar]
  147. Wang R, Ma J, Wu Q, Xia J, Miele L, Sarkar FH, Wang Z, 2013. Functional role of miR-34 family in human cancer. Curr. Drug Targets 14, 1185–1191. [DOI] [PubMed] [Google Scholar]
  148. Wang L, He J, Xu H, Xu L, Li N, 2016. MiR-143 targets CTGF and exerts tumor-suppressing functions in epithelial ovarian cancer. Am. J. Transl. Res 8, 2716–2726. [PMC free article] [PubMed] [Google Scholar]
  149. Wei S, Wei WY, Sedivy JM, 1999. Expression of catalytically active telomerase does not prevent premature senescence caused by overexpression of oncogenic Ha-Ras in normal human fibroblasts. Cancer Res 59, 1539–1543. [PubMed] [Google Scholar]
  150. Wong P, Iwasaki M, Somervaille TC, Ficara F, Carico C, Arnold C, Chen CZ, Cleary ML, 2010. The miR-17–92 microRNA polycistron regulates MLL leukemia stem cell potential by modulating p21 expression. Cancer Res 70, 3833–3842. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Wu S, Liu F, Xie L, Peng Y, Lv X, Zhu Y, Zhang Z, He X, 2015. miR-125b suppresses proliferation and invasion by targeting MCL1 in gastric cancer. BioMed Res. Int 2015, 365273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Xu D, Takeshita F, Hino Y, Fukunaga S, Kudo Y, Tamaki A, Matsunaga J, Takahashi RU, Takata T, Shimamoto A, Ochiya T, Tahara H, 2011. miR-22 represses cancer progression by inducing cellular senescence. J. Cell Biol 193, 409–424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Xu N, Zhang L, Meisgen F, Harada M, Heilborn J, Homey B, Grander D, Stahle M, Sonkoly E, Pivarcsi A, 2012. MicroRNA-125b down-regulates matrix metallopeptidase 13 and inhibits cutaneous squamous cell carcinoma cell proliferation, migration, and invasion. J. Biol. Chem 287, 29899–29908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Xu Z, Wang X, Fang L, 2013. MicroRNA miR-24 enhances tumor angiogenesis and metastasis in breast cancer. Cancer Res 73, P4-07-19. [Google Scholar]
  155. Xu F, Pang L, Cai X, Liu X, Yuan S, Fan X, Jiang B, Zhang X, Dou Y, Gorospe M, Wang W, 2014. let-7-repressesed Shc translation delays replicative senescence. Aging Cell 13, 185–192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Xu X, Chen W, Miao R, Zhou Y, Wang Z, Zhang L, Wan Y, Dong Y, Qu K, Liu C, 2015. miR-34a induces cellular senescence via modulation of telomerase activity in human hepatocellular carcinoma by targeting FoxM1/c-Myc pathway. Oncotarget 6, 3988–4004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Xue J, Chi Y, Chen Y, Huang S, Ye X, Niu J, Wang W, Pfeffer LM, Shao ZM, Wu ZH, Wu J, 2016. MiRNA-621 sensitizes breast cancer to chemotherapy by suppressing FBXO11 and enhancing p53 activity. Oncogene 35, 448–458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Yamakuchi M, 2012. MicroRNA regulation of SIRT1. Front. Physiol 3, 68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Yao S, 2016. MicroRNA biogenesis and their functions in regulating stem cell potency and differentiation. Biol. Proced. Online 18, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Ye J, Yao Y, Song Q, Li S, Hu Z, Yu Y, Hu C, Da X, Li H, Chen Q, Wang QK, 2016. Up-regulation of miR-95–3p in hepatocellular carcinoma promotes tumorigenesis by targeting p21 expression. Sci. Rep 6, 34034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Yi C, Wang Q, Wang L, Huang Y, Li L, Liu L, Zhou X, Xie G, Kang T, Wang H, Zeng M, Ma J, Zeng Y, Yun JP, 2012. MiR-663, a microRNA targeting p21(WAF1/CIP1), promotes the proliferation and tumorigenesis of nasopharyngeal carcinoma. Oncogene 31, 4421–4433. [DOI] [PubMed] [Google Scholar]
  162. Zhai L, Ma C, Li W, Yang S, Liu Z, 2016. miR-143 suppresses epithelial-mesenchymal transition and inhibits tumor growth of breast cancer through down-regulation of ERK5. Mol. Carcinog 55, 1990–2000. [DOI] [PubMed] [Google Scholar]
  163. Zhang RG, Poustovoitov MV, Ye XF, Santos HA, Chen W, Daganzo SM, Erzberger JP, Serebriiskii IG, Canutescu AA, Dunbrack RL, Pehrson JR, Berger JM, Kaufman PD, Adams PD, 2005. Formation of MacroH2A-containing senescence-associated heterochromatin foci and senescence driven by ASF1a and HIRA. Dev. Cell 8, 19–30. [DOI] [PubMed] [Google Scholar]
  164. Zhang RG, Chen W, Adams PD, 2007. Molecular dissection of formation of senescence-associated heterochromatin foci. Mol. Cell. Biol 27, 2343–2358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Zhang Y, Wang Y, Wang X, Zhang Y, Eisner GM, Asico LD, Jose PA, Zeng C, 2011. Insulin promotes vascular smooth muscle cell proliferation via microRNA-208-mediated downregulation of p21. J. Hypertens 29, 1560–1568. [DOI] [PubMed] [Google Scholar]
  166. Zhang D, Song Y, Wang Y, Liu X, Liu X, Ma X, 2016a. Insight of in vitro small-interfering RNA release from chitosan nanoparticles under enzymolysis with forster resonance energy transfer analysis. J. Pharm. Sci 105, 301–307. [DOI] [PubMed] [Google Scholar]
  167. Zhang HY, Duan JJ, Qu YJ, Deng T, Liu R, Zhang L, Bai M, Li JL, Ning T, Ge SH, Wang X, Wang ZZ, Fan Q, Li HL, Ying GG, Huang DZ, Ba Y, 2016b. Onco-miR-24 regulates cell growth and apoptosis by targeting BCL2L11 in gastric cancer. Protein Cell 7, 141–151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Zhang J, Jia J, Zhao L, Li X, Xie Q, Chen X, Wang J, Lu F, 2016c. Down-regulation of microRNA-9 leads to activation of IL-6/Jak/STAT3 pathway through directly targeting IL-6 in HeLa cell. Mol. Carcinog 55, 732–742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Zhang M, Gong W, Zuo B, Chu B, Tang Z, Zhang Y, Yang Y, Zhou D, Weng M, Qin Y, Ma M, Jiang A, Ma F, Quan Z, 2016d. The microRNA miR-33a suppresses IL-6-induced tumor progression by binding twist in gallbladder cancer. Oncotarget 7, 78640–78652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Zhang Y, Guo M, Xin N, Shao Z, Zhang X, Zhang Y, Chen J, Zheng S, Fu L, Wang Y, Zhou D, Chen H, Huang Y, Dong R, Xiao C, Liu Y, Geng D, 2016e. Decreased microRNA miR-181c expression in peripheral blood mononuclear cells correlates with elevated serum levels of IL-7 and IL-17 in patients with myasthenia gravis. Clin. Exp. Med 16, 413–421. [DOI] [PubMed] [Google Scholar]
  171. Zhao GB, Liu LJ, Zhao TS, Jin SD, Jiang SB, Cao SQ, Han JQ, Xin YZ, Dong Q, Liu X, Cui J, 2015. Upregulation of miR-24 promotes cell proliferation by targeting NAIF1 in non-small cell lung cancer. Tumor Biol 36, 3693–3701. [DOI] [PubMed] [Google Scholar]
  172. Zheng X, Chopp M, Lu Y, Buller B, Jiang F, 2013. MiR-15b and miR-152 reduce glioma cell invasion and angiogenesis via NRP-2 and MMP-3. Cancer Lett 329, 146–154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Zheng L, Zhang Y, Liu Y, Zhou M, Lu Y, Yuan L, Zhang C, Hong M, Wang S, Li X, 2015. MiR-106b induces cell radioresistance via the PTEN/PI3K/AKT pathways and p21 in colorectal cancer. J. Transl. Med 13, 252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Zhou Z, Yin Y, Chang Q, Sun G, Lin J, Dai Y, 2016. Downregulation of B-myb promotes senescence via the ROS-mediated p53/p21 pathway, in vascular endothelial cells. Cell Prolif 50, e12319. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES