Open Access

Mechanisms shaping the role of ERK1/2 in cellular senescence (Review)

  • Authors:
    • Junrong Zou
    • Tingting Lei
    • Pei Guo
    • Jason Yu
    • Qichao Xu
    • Yunfei Luo
    • Rong Ke
    • Deqiang Huang
  • View Affiliations

  • Published online on: November 29, 2018     https://doi.org/10.3892/mmr.2018.9712
  • Pages: 759-770
  • Copyright: © Zou et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Senescence is a result of cellular stress and is a potential mechanism for regulating cancer. As a member of the mitogen‑activated protein kinase family, ERK1/2 (extracellular signal‑regulated protein kinase) has an important role in delivering extracellular signals to the nucleus, and these signals regulate the cell cycle, cell proliferation and cell development. Previous studies demonstrated that ERK1/2 is closely associated with cell aging; however other previous studies suggested that ERK1/2 exerts an opposite effect on aging models and target proteins, even within the same cell model. Recent studies demonstrated that the effect of ERK1/2 on aging is likely associated with its target proteins and regulators, negative feedback loops, phosphorylated ERK1/2 factors and ERK1/2 translocation from the cytoplasm to the nucleus. The present review aims to examine the mechanism of ERK1/2 and discuss its role in cellular outcomes and novel drug development.

Introduction

The Hayflick limit is a response to cellular lesions, which are triggered by multiple mechanisms, including replicative senescence, oncogene activation, telomerase dysfunction and DNA lesions (13). Senescent cells arrested at the G1 phase demonstrated more properties associated with dysfunctional cells compared with normal cells (4). Although senescence is an undesirable stress for normal cells, it is beneficial for the body as it restrains excessive proliferation of tumor cells. Therefore, senescence is used as a means of suppressing cancer and is an important cancer treatment method (58).

Extracellular signal-regulated protein kinase (ERK)1/2 is a mitogen-activated protein kinase (MAPK) family protein with typical cascade signaling characteristics and serves an important role in signal transduction pathways and the function of transcription factors, including activator protein-1, proto-oncogene c-Fos (c-Fos) and ETS domain-containing protein Elk-1 (Elk1) (9). The majority of research has focused on its regulatory effect on cell growth and differentiation (1014); however, a number of previous studies demonstrated that ERK1/2 promotes cell senescence (1517). Based on these characteristics, numerous small molecule MAPK/ERK kinase (MEK) inhibitors were examined in early-phase clinical trials, including PD098059, U0126, CI-1040, PD0325901 and AZD6244 (18); however, none of them were approved by The Food and Drug Administration due to adverse side effects or other toxic reactions (18). Many of these inhibitors negatively affected normal and abnormal cells. Notably, these effects may have been the result of the dual roles of ERK1/2 in senescence, as demonstrated by other previous studies (19,20). The present review examines the mechanisms regulating the role of ERK1/2 in cell senescence and suggests that ERK1/2 is a potentially useful target in treating cancer.

Senescence, ERK1/2 and cell fate

ERK1/2 signaling pathway

The mammalian MAPKs consist of cytoplasmic serine/threonine kinases that are involved in the transduction of signals from the surface to the interior of the cell. This family includes the ERK family (ERK1-8), the p38 kinase family (p38 α/β/γ/δ) and the c-Jun N-terminal kinase family (JNK1-3, additionally termed stress-activated protein kinase) (9). With a number of substrate docking and enzyme recruitment sites (21), ERK1/2 (MAPK1/3) is a multifunctional serine/threonine kinase that is able to phosphorylate numerous substrates, including protein kinases, signal effectors, receptors, cellular scaffold proteins and nuclear transcriptional regulators (21). At present, five types of ERK isoforms are known. ERK1 and ERK2 are thought to be the most important isoforms with 84% homology for the primary sequences and similar functions (21).

The Ras/Raf/MEK/ERK1/2 signaling pathway is a small GTPase ligation of activated tyrosine receptors and cytoplasmic kinase signal transduction cascades. The key point of activation is to transmit a signal from tyrosine receptors, including epidermal growth factor receptor (EGFR), which subsequently recruit Son of sevenless (SOS) through intracellular Shc and Grb2 domains, ultimately catalyzing the conversion of inactive Ras/guanosine diphosphate to the active Ras/guanosine triphosphate complex (22). As an activator of ERK1/2, MEK1/2 catalyzes the phosphorylation of ERK1/2 at Tyr204/187 and Thr202/185 by casein kinase 2 (CK2) (23). This enzyme subsequently binds to importin7 and translocates ERK1/2 from the cytoplasm to the nucleus (24), where it functions as an upstream regulator of substrate genes that encode for transcription factors, including Elk1, c-Myc, signal transducers and activators of transcription (STATs), c-Jun and c-Fos (21). These transcription factors regulate their counterpart target genes to alter the expression or activity of various proteins and are involved in the regulation of a large variety of processes, including adhesion, cell cycle progression, migration, survival, differentiation, metabolism, proliferation and transcription (21). The Ras/Raf/MEK/ERK cascade is a highly efficient signaling pathway, aided by scaffold proteins, including kinase suppressor of Ras, MEK partner 1/p14 complex, β-arrestins, fibroblast growth factor receptor substrate 2, MAPK organizer 1 and flotillin-1 (25). The function of scaffold proteins is characterized by combinatorial inhibition, which is the stoichiometry of a scaffold and its signaling partners; the expression levels of scaffold proteins should not be too high (the kinase and its substrate may each bind to an individual scaffold protein) or too low (the phosphorylation of the cascade is sub-optimal) (26). With these different scaffold proteins, the phosphorylation of different isoforms is accurately regulated; the scaffold protein MEK protein 1 specifically binds ERK1, not ERK2 (27).

Role of ERK1/2 in cellular senescence

Cellular senescence was first observed in cultured fibrocytes, when the Hayflick limit demonstrated that as the cells divided, their cell cycle became arrested (replicative senescence) (28). Senescent cells have abnormal metabolic activity (3), accompanied by morphological, biological and genetic alterations. When β-galactosidase expression (an important senescence marker) increases (29), the cell cycle is arrested at the G1/S phase (30). Cell cycle dependent kinase (CDK) and cyclin A activity additionally decrease with increased activity of cyclin-dependent kinase inhibitors (p16INK4a and p21) (31). Without the protection of histones, mitochondrial reactive oxygen species (ROS) may damage mitochondrial DNA, which induces a series of oxidative stress reactions (32). Telomere shortening, which causes DNA to lose protection from the telomeres, is another feature of senescence and leads to DNA integration and degradation. Oxidative stress reactions and mitochondrial dysfunction (33) accelerate the shortening of telomeres. Cellular senescence is triggered by a number of signaling pathways and mechanisms, including DNA injury, telomerase dysfunction, oncogenes, oxidative stress reactions and mitochondrial dysfunction (34).

ERK1/2 is an important messenger for extracellular and intracellular signals, which serve a vital role in processes, including proliferation, differentiation, cytoskeleton construction and cellular senescence (35). In the majority of cases, ERK1/2 is a regulator of cellular proliferation; however, it has been identified that ERK1/2 may additionally promote senescence. Strong, constitutively active expression of MEK1 (an upstream activator of ERK1/2) in non-immortalized intestinal epithelial cells (HIEC cells) promotes cellular senescence, whereas, in immortalized intestinal cells (IEC-6 cells), it does not, suggesting that cell type may serve a role (36). This phenomenon requires further investigation to improve the clinical use of ERK1/2-associated reagents, including ERK1/2 inhibitors.

ERK1/2 promotes cellular proliferation

ERK1/2 is associated with cell survival, proliferation and development. To investigate the role of ERK1/2 in different cell types and animal models, a number of previous studies investigating this were reviewed. ERK1/2 functions more frequently as a cellular proliferation marker than as a dual role kinase (Table I) (3754). Based on these previous studies, cellular proliferation is primarily regulated by the effects of ERK1/2 on cell cycle entry and protein synthesis.

Table I.

Models of ERK controlled proliferation.

Table I.

Models of ERK controlled proliferation.

Author, yearCellular modeCell and tissue typeKey signaling pathway and mechanism(Refs.)
Ravasi et al, 2006Human bronchi isolatedHuman airway smooth primary cell LTD4-EGFR-Ras-ERK1/2 muscle cells(37)
Gong et al, 2006Mouse isolated primary cellNeural progenitor cellSDF1-ERK1/2 or PI3K(38)
Iyengar et al, 2006Lens epithelial explantsLens epithelial explantsAKT-ERK1/2(39)
Wang et al, 2006Oral tongue squamous cell carcinomasOral tongue carcinomasERK1/2-Cyclin D1(40)
De Rosa et al, 2007Human primary isolated Treg cellsHuman Treg cellsLeptin inhibits ERK1/2(41)a
Li et al, 2007Panc-1, Panc-03.27, ASPC-1 and PL45Human pancreatic cancer cellsThya1-ERK1/2 or JNK(42)
Li et al, 2008AGS, RGM-1 cellRat gastric mucosal cellsFas Ag-FasL coupling activation of ERK1/2(43)
He et al, 2008C18-4 cellsMouse spermatogonial stem cell GNDF-shc-Grb-Ras-ERK1/2-CREB1-C-fos-cyclinA/CDK2(44)
Mancinelli et al, 2009Cholangiocyte and animal modelMouse cholangiocyte FSH-cAMP-ERK1/2(45)
Sirianni et al, 2010MCF-7Human breast cancer cellNo data(46)
Yang and Han, 2010Mouse isolated primary cellMouse immature Sertoli cells in testis tissue GDNF-NCAM-ERK1/2(47)
Lee and Kay, 2011Rabbit isolated primary cellRabbit corneal endothelial cells FGF2-ERK1/2-Cdc25A-CDK2(48)
Gao et al, 2014L02Human hepatocyte HPS/Src/EGFR/ERK(49)
Tocharus et al, 2014Rat isolated primary cellNeural stem/progenitor cellsMelatonin receptor/c-Raf/ERK1/2(50)
Wu et al, 2014NHEKsHuman epidermal keratinocytesEGFR/ERK/AQP3(51)
Liu et al, 2015A549, SPCA1 and PC9 cellsHuman non-small cell lung cancer PTP1B-src-ERK1/2(52)
Wang et al, 2016Mouse isolated primary Müller cellMouse retinal Müller cellsNGF/TrkA-VEGF ERK1/2(53)
Kim et al, 2017Human ASM cellsHuman airway smooth muscle cellsNo data(54)

a Study mentioned dual role of ERK1/2.

DNA and protein synthesis

As a transcription factor regulator, ERK1/2 transduces signals from the cell membrane to the nucleus. ERK1/2 may additionally regulate carbamoyl phosphate synthetase II (55), which catalyzes the initial rate-limiting step in the de novo synthesis of pyrimidine nucleotides. Furthermore, it was identified that ERK1/2 phosphorylates high motility group boxes of nucleolar transcription factor 1, an RNA polymerase I factor transcriptional enhancer that enhances ribosomal RNA genes (56).

Mendoza et al (57) demonstrated that the MEK1/2-ERK1/2 pathway cross-talks with the phosphoinositide 3 kinase (PI3K)/protein kinase B (AKT)/mammalian target of rapamycin (mTOR) pathway via cross-inhibition, cross-activation and pathway convergence on substrates. In mouse uterine epithelial cells, activation of protein kinase C by estradiol-17β promotes protein synthesis by activating the ERK-mTOR-40s ribosomal protein S6 cascade (58). ERK1 and ERK2 enhance protein translation by increasing the ability of eukaryotic translation factor-4E (59) to recruit protein-synthesizing ribosomes and other protein synthesis initiation factors to the mRNA. This recruitment includes nuclear substrates (ternary complex factor, Elk1, and c-Fos), cytoplasmic substrates [40s ribosomal protein S6 kinase (RSK) family], cytoskeletal proteins and proteins of the nuclear pore complex, many of which serve a direct role in cellular proliferation and development (21). A previous study additionally demonstrated that ERK1/2 may functionally dephosphorylate the tuberous sclerosis 1 and 2 proteins (TSC1/2) complex via its downstream RSK in HEK293 cells (60).

Cell cycle entry

ERK1/2 is involved in G1/S and G2/M transitions (61). During G1/S, ERK1/2 regulates cyclin D1 transcription through the Fos family of proteins (62) and Myc (63,64). In G2/M, ERK1/2 is involved in the nuclear translocation of cyclin B1 by phosphorylating two of four sites within the cytoplasmic retention sequence of cyclin B1 (65) and inhibiting the negative phosphorylation of cell division control protein 2 homolog by myelin transcription factor 1 via RSK2 (66).

In addition to the substrates involved in cell proliferation, ERK1/2 additionally regulates cellular tumor antigen p53 (p53) phosphorylation. p53 is a tumor suppressor protein and functions as a transcription factor by binding to a number of genes, including cyclin-dependent kinase inhibitor 1A, which encodes p21. p21 binds and inactivates CDKs, which are crucial for cell entry into the G1/S phase (67,68). The association between ERK1/2 and p53 remains unclear. A previous study suggested that p53 functions upstream of ERK1/2 (69); however, the most widely accepted hypothesis is that ERK1/2 regulates p53 by activating STAT3 (70) and other transcription factors. ERK1/2 and p53 have hundreds of substrates, thus, it is easy for them to engage in crosstalk, as is the case with dual-specificity phosphatases (DUSPs) (71). The effect of ERK1/2 on its downstream substrates may accelerate the degradation of p53 (72). ERK1/2 regulates p53 phosphorylation [a form that protects p53 from E3 ubiquitin-protein ligase Mdm2 (73)] through the forkhead box M1/c-myc/polycomb complex protein BMI-1 pathway, which inhibits p19 phosphorylation, attenuating cellular senescence (74).

ERK1/2 regulates mitochondria

Mitochondria not only provide energy to cells; however, additionally serve a decisive role in cell fate. Mitochondria within the respiratory chain are responsible for maintaining the proton gradient and providing various respiratory enzymes; it was demonstrated that the proton gradient is not just associated with adenosine triphosphate synthesis. Rasola et al (75) identified that ERK1/2 phosphorylates glycogen synthase kinase-3β, inhibiting permeability transition pore opening by regulating cyclophilin D and preventing the release of apoptotic substances, including mitochondrial cytochrome C, ROS, Ca2+ and free radicals. The number of mitochondria is additionally an important hallmark of cellular proliferation. A previous study demonstrated that ERK2 may phosphorylate dynamin-1-like protein (an important regulator of mitochondrial fission) on serine 616 in several tumor models (76), resulting in tumor growth.

ERK1/2 promotes cellular senescence through several mechanisms

Based on extensive investigations in a variety of cell types, previous studies identified that ERK1/2 may additionally facilitate cellular senescence under certain circumstances. In the present review, a number of previous studies are discussed to gain a better understanding of the role of ERK1/2 during cellular senescence and the underlying mechanisms behind its control. In contrast, the role of ERK1/2 in cellular proliferation was only studied in numerous cell types, primarily fibroblasts, providing limited information. The previous studies investigating ERK1/2 involvement in cellular senescence (Table II) (7790) identified a number of possible mechanisms for the role of ERK1/2 that are associated with abnormal signaling of negative feedback loops, caused by constitutive and overexpressed ERK1/2 (20,77,78), and ERK1/2 cellular localization (79).

Table II.

Models of ERK-induced senescence.

Table II.

Models of ERK-induced senescence.

Author, yearCellular modeCell and tissue typeKey signaling pathway and mechanism(Refs.)
Wang et al, 2002BJ human foreskin fibroblastsHuman fibroblasts ERK-MKK3/6-p38-p53-p16(19)a
Lin et al, 1998IMR90Human fibroblastsConstitutive activation active p53/p16(20,77,78)
Zhu et al, 1998
Cammarano et al, 2005
Kim et al, 2000TIG-3 and TIG-7Human fibroblastsTranslocation from nucleus of p-ERK1/2 which was helped by E1A(79)
Lim et al, 2000Primary isolated and PA137Human and mouse fibroblasts H-rasV12S35/MEK/Racl/RhoA(80)
Chaturvedi et al, 2003KCsHuman neonatal foreskinRas/Raf/ERK elevate p16 and p14ARF(81)
Kim et al, 2003Human diploid fibroblastsHuman fibroblastsROS inactivation PPP1/2A (resensitise Raf-1), lead to ERK1/2 constitutive activation(82)
Todd et al, 2004CC139, RAT-1, 3T3L1Human/mouse fibroblast and preadipocytep38 and ERK1/2 cooperate to induce p21CIP1 expression(83)
Klein et al, 2005A549Human non-small cell lung cancer cellSustained activation of ERK1/2 by discodermolide(84)
Albrecht et al, 2007PC-3Prostate cancer cellEGCG (drug) activate ERK1/2 via MEK1/2 independently(85)
Deschênes-Simard et al, 2013293T, Hela, IMR90Human embryonic kidney cell, etc.Sustained activation of ERK1/2 induce senescence associated protein degradation(86)
Zhu et al, 2014Primary keratinocytes, HaCat cellPrimary mouse kerationcytes and human skin cellPPARβ/δ increases ERK1/2 activation, which upregulate expression of p53 and p21(87)
Wang et al, 2013HCT116Human epithelial cellsASPP2 stimulate HRas induced ERK1/2 activation(88)
El Bezawy et al, 2017MesoII, STO, MP115, MP4 and MP8 cellsHuman diffuse malignant peritoneal mesotheliomamiR-34a induced persistent activation of ERK1/2(89)
Del Nogal et al, 2014Human glomerular mesangial cellsHuman primary renal glomerular cellConstitutive activation of Ras, and ROS production elevated(90)

a Study mentioned dual role of ERK1/2.

Crosstalk and negative feedback loops associated with ERK1/2-induced cellular senescence
Regulation of the MAPK signaling pathway

A number of previous studies demonstrated that negative regulation of ERK1/2 within MAPK signaling cascades regulate ERK1/2 signaling. ERK1/2 phosphorylates proteins within this cascade at alternate sites, which interrupts the normal binding behavior of their respective downstream substrates (91). ERK1/2 phosphorylates EGFR at T669 (92) and decreases constitutive tyrosine phosphorylation activity, decreasing the ability of the phosphorylated loop to cross-activate other adaptors (93). ERK1/2 may additionally phosphorylate dual specificity Cdc25C at T48, which dephosphorylates EGFR at Y1068 (94). Furthermore, ERK1/2 was identified to phosphorylate MAPK signaling components, including fibroblast growth factor receptor (FGFR) at S777 (95); SOS1 at S1132, S1167, S1178 and S1193 (96); fibroblast growth factor receptor substrate 2 (FRS2)α at T132, T135, T138, T376, T452, T455, T458 and T463 (97,98); RAF proto-oncogene serine/threonine-protein kinase (Raf-1) at S29, S289, S296, S301 and S642 (99103); serine/threonine-protein kinase B-raf (B-Raf) at S151, T401, S750, T753 and S642 (104,105); MEK1 at T292 and T386 (106,107); and kinesin suppressor or Ras 1 at T260, T274, S320, S443 and S463 (108,109) (Fig. 1). The phosphorylation of all these components results in a disruption of binding to downstream substrates.

In other previous study, Fey et al (110) developed a dynamic model of the multiple MAPK cascade interactions and feedback systems of specific proteins of the MAPK pathway using mathematical analysis. Based on their model, p38 inhibits ERK through Ser/Thr protein phosphatase-2A, and JNK restrains p38 and ERK through induction of DUSPs (JNK upregulates transcription of DUSP1; Fig. 1).

Dephosphorylation of ERK1/2

ERK1/2 requires dual phosphorylation of threonine and tyrosine residues to acquire its biological kinase function. Dual-specificity Thr/Tyr phosphatases [DUSPs; additionally termed MAPK phosphatases (MKP)] represent a large family that regulates the activity of MAPKs by dephosphorylating threonine and tyrosine residues within the activation loop of MAPKs, which in turn regulates the biologically active form of ERK1/2 in the cytoplasm and the nucleus (111). DUSP1, DUSP2, DUSP4 and DUSP5 are located in the nucleus, whereas DUSP6, DUSP7 and DUSP9 are located in the cytoplasm (112). Their binding to ERK1/2 is regulated by a conserved motif within the amino-terminal non-catalytic domain (kinase-interacting motif) of the protein (113,114) and results in a significant increase in catalytic activity, which deprives ERK1/2 of the phosphate group. The interaction between MAPKs and DUSPs is a two-way regulation; MAPKs are able to upregulate transcription of DUSPs (113), primarily those in the cytoplasm (115), in a delayed manner following MAPK activation, whereas, DUSPs strictly regulate MAPK signaling.

Sprouty [protein sprout homolog (Spry) 1–4] is another ERK1/2 regulator family that is not well-studied. A previous Japanese study demonstrated that Spry1 and Spry2 are phosphorylated at Y53 and Y55, which creates a docking site for growth factor receptor-bound protein 2 at the Src homology 2 domain and consequently disrupts association with the FGFR adaptor FRS2 in C2C12 cells (116). Other previous studies suggested that the spry protein inhibits receptor tyrosine kinase signaling, which suppresses the activation of Ras in BRAFV600E melanomas (117,118). Another previous study demonstrated that in 293T cells, Spry1 and Spry2 regulate Raf-1 by directly binding to it (119). Lake et al (91) suggested that Sprouty serves its functions at multiple nodes in a context-specific manner.

Crosstalk between Ras/ERK1/2 and PI3K/AKT signaling

In contrast to signaling that regulates cell proliferation, AKT/mTOR and ERK1/2 engage in crosstalk that sustains cell proliferation and survival, which in turn helps cells escape from either PI3K/AKT or ERK1/2 suppression (120). Crystal structure analysis demonstrated that astrocytic phosphoprotein PEA-15 (PEA-15) may efficiently bind to the ERK2 activation loop at the Thr-X-Tyr region (121), activating transportation of ERK1/2 from the nucleus. Sinha et al (122) observed that ERK1/2 decreases phosphorylated (p)-AKT expression levels in mouse renal proximal tubular cells via Ras/PI3K through a negative feedback pathway, whereas AKT phosphorylates and stabilizes PEA-15, which subsequently decreases the nuclear localization of ERK1/2 (123) and induces cellular senescence. Although the regulatory nodes shared by Ras/MAPK and PI3K/AKT signaling are complicated, the two signaling pathways are influenced by co-effectors, including TSC1/2, mTOR, ER and S6, and regulate each other in concentration- and context-dependent manners (124).

Crosstalk between ERK1/2 and p53

As a messenger of extracellular and intracellular proteins, ERK1/2 has many substrates. MEK1/2 functions upstream of ERK1/2 and serves a vital regulatory role in ERK1/2 activation. MEK1/ERK signaling promotes cell proliferation, whereas MEK2/ERK signaling promotes G1/S cell cycle arrest (125). In different tissues and cell types, including mouse embryos and fibroblasts, cellular senescence induced by the Ras/Raf/MEK/ERK signaling pathway is dependent on the integrity of p16/INK4A, p21 and p53; in human primary fibroblasts, inhibition of either p16 or p53 is not able to reverse ERK1/2-induced senescence (17,20,78). However, a previous study demonstrated that the matrix cell protein G1/S-specific cyclin CCN1 induces senescence through the p53/p21 pathway and inhibits lung cancer growth (36). p53 and ERK1/2 have two-way regulation, which means there is a negative feedback loop between ERK1/2 and p53. Lee et al (126) suggested that a novel p53 target protein, Raf kinase inhibitor protein, inhibits ERK1/2 by affecting Raf proteins and promoting senescence. Notably, p53 may regulate the transcription of all nuclear DUSPs (DUSP1/2/4/5) (127129).

Chemicals or gene mutations lead to ERK1/2-associated senescence

Exposure of cells to certain bioactive chemicals, including the natural ethanolic Rhus coriaria extract may lead to activation of ERK1/2 and p21 upregulation (130). The microtubule stabilizing agent discodermolide was identified to induce cellular senescence due to overexpressed ERK1/2 in A549 cells (66). Administration of epigallocatechin-3-gallate leads to cellular senescence during PC-3 prostate cancer cell proliferation via MEK-independent ERK1/2 activation (85). In addition, a Ras or B-Raf mutation is common in tumors, particularly in malignant tumors, which may lead to sustained activation of ERK1/2 signaling. In specific cases, downstream proteins are activated by ERK1/2 aberrantly, resulting in different physiological effects (103). Expression of proteins downstream of ERK1/2, including BRAF-induced insulin-like growth factor-binding protein 7 (IGFBP7) in normal melanoma cells, is low and primarily controlled by autocrine or paracrine functions that influence cell proliferation (103). Examples of such phenomenon include BRAFV600E-positive nevi, which contain high BRAF expression; the continuously activated RAF/MEK/ERK pathway, which increases IGFBP7 expression; and high expression levels of IGFBP7, which inhibit the RAF/MEK/ERK pathway within cells. However, in melanoma cells, IGFBP7 expression is lost, which results in uncontrolled proliferation (131).

Relationship between ERK1/2 and micro (mi)RNAs

miRNAs are important for regulating cell biology. Previous studies demonstrated that miRNA-34a induces persistent activation of ERK1/2, leading to cellular senescence via inhibition of p53 signaling (89) and MEK1/2 (132). miR-21 increases the expression level of p-ERK1/2 by inhibiting Spry1 (116) and Spry2 (133).

Cellular localization and duration influence ERK1/2-associated cell fate
Negative feedback occurs when ERK1/2 is overexpressed

A number of previous studies suggest that ERK1/2-induced cellular senescence may be associated with the strength of ERK1/2 signals and the duration of its activation (20,77,78,86,90,134). As ERK1/2 has many substrates, different biological effects may occur when the signal proteins compete for the same target protein. As discussed above, many negative feedback loops and regulatory nodes shape ERK1/2 signaling, and when ERK1/2 is overexpressed, it may activate those negative loops. However, this regulatory system is dependent on the cellular context.

Translocation from the cytoplasm to the nucleus determines the role of ERK1/2

Scaffold proteins facilitate protein translocation, and translocation of ERK1/2 from the cytoplasm to the nucleus is essential for regulation of the cell cycle and cellular proliferation (135). This process requires dual-phosphorylation of specific residues within the activation loop. ERK is initially phosphorylated by MAPK/ERK kinase (MEK) on the Thr-Glu-Tyr motif, with subsequent phosphorylation on the Ser-244/Pro-245/Ser-246 (SPS) nuclear translocation sequence (NTS) (108,109). This is achieved primarily by CK2 to generate pSPS-pERK, which binds to the shuttling protein importin7 (Imp7) (136,137). A previous study investigating mouse embryo fibroblasts demonstrated that Tm5NM1-containing actin filaments facilitate the binding of pSPS-pERK (Ser-244/Pro-245/Ser-246) and Imp7, possibly by functioning as a scaffold and/or recruiting myosin motors to assist in the physical transportation of pSPS-pERK from the cytoplasm to the nucleus (136). Inhibiting the binding of pSPS-pERK and Imp7 appears to be an effective way of blocking the translocation of dual-phosphorylated ERK. Plotnikov et al (135) developed a myristoylated, NTS-derived phosphomimetic peptide (EPE peptide) that competes with the binding of Imp7 and blocks this process in a number of cell lines.

ETS translocation variant 4 (E1A) is additionally a negative regulator of activated ERK1/2 translocation from the cytoplasm to the nucleus (79,138). A previous study examining normal human diploid fibroblast IMR90 cells demonstrated that E1A decreases expression levels of PEA-15 (139), an ERK1/2 nuclear export factor, and increases expression levels of MKP1/DUSP1 and DUSP5 (138). The translocation regulation model is presented in Fig. 2.

Conclusion

ERK1/2 serves a vital role in cellular outcomes, which involve numerous substrates, regulators and scaffolding proteins. If a small dose of a MEK1/2 inhibitor, including U0126 or PD098059, is applied to a cell, the cell will recover by utilizing a compensatory pathway to regain its proliferation capacity. Even in the same type of cell or animal model, ERK1/2 serves a dual role in cellular senescence under different circumstances, which are dose- and duration-dependent. Cells subjected to drugs against ERK1/2 may additionally contain mutations in upstream (AKT, MEK and Ras/Raf) or downstream (p21 and p53) proteins. Analyzing cells for Ras/Raf/MAPK mutations and testing cell sensitivity may help to determine the proper dosage and duration of drug administration. In terms of drug development, altering the translocation of ERK1/2 from the cytoplasm to the nucleus (135), which is primarily required for induction of cell proliferation, may help to decrease the proliferation of cancer cells. However, as cells may shift their proliferation signals between a number of different proteins, a combination of numerous anti-proliferation tactics require consideration.

Acknowledgements

Not applicable.

Funding

The present study was supported by the National Natural Science Foundation of China (grant nos. 81460374, 31460304 and 81572753), the Natural Science Foundation of Jiangxi Province of China (grant nos. 20171BAB205062 and 20171BCB23086) and the Education Department of Jiangxi Province of China (grant no. 160032).

Availability of data and materials

Not applicable.

Authors' contributions

JZ and TL contributed equally in reviewing the publications and writing this manuscript. PG and DH drafted the manuscript. JY, QX, YL and RK designed the schema of ERK1/2 signaling. All authors read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Cristofalo VJ, Lorenzini A, Allen RG, Torres C and Tresini M: Replicative senescence: A critical review. Mech Ageing Dev. 125:827–848. 2004. View Article : Google Scholar : PubMed/NCBI

2 

Zeiser R: Trametinib. Recent Results Cancer Res. 201:241–248. 2014. View Article : Google Scholar : PubMed/NCBI

3 

Salama R, Sadaie M, Hoare M and Narita M: Cellular senescence and its effector programs. Genes Dev. 28:99–114. 2014. View Article : Google Scholar : PubMed/NCBI

4 

Tominaga K: The emerging role of senescent cells in tissue homeostasis and pathophysiology. Pathobiol Aging Age Relat Dis. 5:277432015. View Article : Google Scholar : PubMed/NCBI

5 

Gewirtz DA: Autophagy and senescence in cancer therapy. J Cell Physiol. 229:6–9. 2014.PubMed/NCBI

6 

Ohtani N, Mann DJ and Hara E: Cellular senescence: Its role in tumor suppression and aging. Cancer Sci. 100:792–797. 2009. View Article : Google Scholar : PubMed/NCBI

7 

Wu CH, van Riggelen J, Yetil A, Fan AC, Bachireddy P and Felsher DW: Cellular senescence is an important mechanism of tumor regression upon c-Myc inactivation. Proc Natl Acad Sci USA. 104:13028–13033. 2007. View Article : Google Scholar : PubMed/NCBI

8 

Roberts PJ and Der CJ: Targeting the Raf-MEK-ERK mitogen-activated protein kinase cascade for the treatment of cancer. Oncogene. 26:3291–3310. 2007. View Article : Google Scholar : PubMed/NCBI

9 

Plotnikov A, Zehorai E, Procaccia S and Seger R: The MAPK cascades: Signaling components, nuclear roles and mechanisms of nuclear translocation. Biochim Biophys Acta. 1813:1619–1633. 2011. View Article : Google Scholar : PubMed/NCBI

10 

Tan X, Wang YL, Yang XL and Zhang DD: Ethyl acetate extract of Artemisia anomala S. Moore displays potent anti-inflammatory effect. Evid Based Complement Alternat Med. 2014:6813522014. View Article : Google Scholar : PubMed/NCBI

11 

Montero-Conde C, Ruiz-Llorente S, Dominguez JM, Knauf JA, Viale A, Sherman EJ, Ryder M, Ghossein RA, Rosen N and Fagin JA: Relief of feedback inhibition of HER3 transcription by RAF and MEK inhibitors attenuates their antitumor effects in BRAF-mutant thyroid carcinomas. Cancer Discov. 3:520–533. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Bell RM, Kunuthur SP, Hendry C, Bruce-Hickman D, Davidson S and Yellon DM: Matrix metalloproteinase inhibition protects CyPD knockout mice independently of RISK/mPTP signalling: A parallel pathway to protection. Basic Res Cardiol. 108:3312013. View Article : Google Scholar : PubMed/NCBI

13 

Wu J, Xu J, Eksioglu EA, Chen X, Zhou J, Fortenbery N, Wei S and Dong J: Icariside II induces apoptosis of melanoma cells through the downregulation of survival pathways. Nutr Cancer. 65:110–117. 2013. View Article : Google Scholar : PubMed/NCBI

14 

Desar IM, Gilles R, van Herpen CM, Timmer-Bonte AJ, Cantarini MV, van der Graaf WT and Oyen WJ: (18)F-FLT-PET for response evaluation of MEK inhibitor selumetinib (AZD6244, ARRY-142886) in patients with solid tumors. World J Nucl Med. 11:65–69. 2012. View Article : Google Scholar : PubMed/NCBI

15 

Barlin JN, Jelinic P, Olvera N, Bogomolniy F, Bisogna M, Dao F, Barakat RR, Chi DS and Levine DA: Validated gene targets associated with curatively treated advanced serous ovarian carcinoma. Gynecol Oncol. 128:512–517. 2013. View Article : Google Scholar : PubMed/NCBI

16 

Rybakova Y, Akkuratov E, Kulebyakin K, Brodskaya O, Dizhevskaya A and Boldyrev A: Receptor-mediated oxidative stress in murine cerebellar neurons is accompanied by phosphorylation of MAP (ERK 1/2) kinase. Curr Aging Sci. 5:225–230. 2012. View Article : Google Scholar : PubMed/NCBI

17 

Zhuang D, Mannava S, Grachtchouk V, Tang WH, Patil S, Wawrzyniak JA, Berman AE, Giordano TJ, Prochownik EV, Soengas MS and Nikiforov MA: C-MYC overexpression is required for continuous suppression of oncogene-induced senescence in melanoma cells. Oncogene. 27:6623–6634. 2008. View Article : Google Scholar : PubMed/NCBI

18 

Wang D, Boerner SA, Winkler JD and LoRusso PM: Clinical experience of MEK inhibitors in cancer therapy. Biochim Biophys Acta. 1773:1248–1255. 2007. View Article : Google Scholar : PubMed/NCBI

19 

Wang W, Chen JX, Liao R, Deng Q, Zhou JJ, Huang S and Sun P: Sequential activation of the MEK-extracellular signal-regulated kinase and MKK3/6-p38 mitogen-activated protein kinase pathways mediates oncogenic ras-induced premature senescence. Mol Cell Biol. 22:3389–3403. 2002. View Article : Google Scholar : PubMed/NCBI

20 

Lin AW, Barradas M, Stone JC, van Aelst L, Serrano M and Lowe SW: Premature senescence involving p53 and p16 is activated in response to constitutive MEK/MAPK mitogenic signaling. Genes Dev. 12:3008–3019. 1998. View Article : Google Scholar : PubMed/NCBI

21 

Roskoski R Jr: ERK1/2 MAP kinases: Structure, function, and regulation. Pharmacol Res. 66:105–143. 2012. View Article : Google Scholar : PubMed/NCBI

22 

Gureasko J, Galush WJ, Boykevisch S, Sondermann H, Bar-Sagi D, Groves JT and Kuriyan J: Membrane-dependent signal integration by the Ras activator Son of sevenless. Nat Struct Mol Biol. 15:452–461. 2008. View Article : Google Scholar : PubMed/NCBI

23 

Chuderland D, Konson A and Seger R: Identification and characterization of a general nuclear translocation signal in signaling proteins. Mol Cell. 31:850–861. 2008. View Article : Google Scholar : PubMed/NCBI

24 

Matsubayashi Y, Fukuda M and Nishida E: Evidence for existence of a nuclear pore complex-mediated, cytosol-independent pathway of nuclear translocation of ERK MAP kinase in permeabilized cells. J Biol Chem. 276:41755–41760. 2001. View Article : Google Scholar : PubMed/NCBI

25 

Meister M, Tomasovic A, Banning A and Tikkanen R: Mitogen-activated protein (MAP) kinase scaffolding proteins: A recount. Int J Mol Sci. 14:4854–4884. 2013. View Article : Google Scholar : PubMed/NCBI

26 

Good MC, Zalatan JG and Lim WA: Scaffold proteins: Hubs for controlling the flow of cellular information. Science. 332:680–686. 2011. View Article : Google Scholar : PubMed/NCBI

27 

Schaeffer HJ, Catling AD, Eblen ST, Collier LS, Krauss A and Weber MJ: MP1: A MEK binding partner that enhances enzymatic activation of the MAP kinase cascade. Science. 281:1668–1671. 1998. View Article : Google Scholar : PubMed/NCBI

28 

Hayflick L and Moorhead PS: The serial cultivation of human diploid cell strains. Exp Cell Res. 25:585–621. 1961. View Article : Google Scholar : PubMed/NCBI

29 

Yang NC and Hu ML: The limitations and validities of senescence associated-beta-galactosidase activity as an aging marker for human foreskin fibroblast Hs68 cells. Exp Gerontol. 40:813–819. 2005. View Article : Google Scholar : PubMed/NCBI

30 

Chen KY: Transcription factors and the down-regulation of G1/S boundary genes in human diploid fibroblasts during senescence. Front Biosci. 2:d417–d426. 1997. View Article : Google Scholar : PubMed/NCBI

31 

Hernandez-Segura A, Nehme J and Demaria M: Hallmarks of cellular senescence. Trends Cell Biol. 28:436–453. 2018. View Article : Google Scholar : PubMed/NCBI

32 

Passos JF and von Zglinicki T: Mitochondria, telomeres and cell senescence. Exp Gerontol. 40:466–472. 2005. View Article : Google Scholar : PubMed/NCBI

33 

Liu L, Trimarchi JR, Smith PJ and Keefe DL: Mitochondrial dysfunction leads to telomere attrition and genomic instability. Aging cell. 1:40–46. 2002. View Article : Google Scholar : PubMed/NCBI

34 

Campisi J: Cellular senescence as a tumor-suppressor mechanism. Trends Cell Biol. 11:S27–S31. 2001. View Article : Google Scholar : PubMed/NCBI

35 

Sun Y, Liu WZ, Liu T, Feng X, Yang N and Zhou HF: Signaling pathway of MAPK/ERK in cell proliferation, differentiation, migration, senescence and apoptosis. J Recept Signal Transduct Res. 35:600–604. 2015. View Article : Google Scholar : PubMed/NCBI

36 

Boucher MJ, Jean D, Vézina A and Rivard N: Dual role of MEK/ERK signaling in senescence and transformation of intestinal epithelial cells. Am J Physiol Gastrointest Liver Physiol. 286:G736–G746. 2004. View Article : Google Scholar : PubMed/NCBI

37 

Ravasi S, Citro S, Viviani B, Capra V and Rovati GE: CysLT1 receptor-induced human airway smooth muscle cells proliferation requires ROS generation, EGF receptor transactivation and ERK1/2 phosphorylation. Respir Res. 7:422006. View Article : Google Scholar : PubMed/NCBI

38 

Gong X, He X, Qi L, Zuo H and Xie Z: Stromal cell derived factor-1 acutely promotes neural progenitor cell proliferation in vitro by a mechanism involving the ERK1/2 and PI-3K signal pathways. Cell Biol Int. 30:466–471. 2006. View Article : Google Scholar : PubMed/NCBI

39 

Iyengar L, Patkunanathan B, Lynch OT, McAvoy JW, Rasko JE and Lovicu FJ: Aqueous humour- and growth factor-induced lens cell proliferation is dependent on MAPK/ERK1/2 and Akt/PI3-K signalling. Exp Eye Res. 83:667–678. 2006. View Article : Google Scholar : PubMed/NCBI

40 

Wang L, Liu T, Nishioka M, Aguirre RL, Win SS and Okada N: Activation of ERK1/2 and cyclin D1 expression in oral tongue squamous cell carcinomas: Relationship between clinicopathological appearances and cell proliferation. Oral Oncol. 42:625–631. 2006. View Article : Google Scholar : PubMed/NCBI

41 

De Rosa V, Procaccini C, Cali G, Pirozzi G, Fontana S, Zappacosta S, La Cava A and Matarese G: A key role of leptin in the control of regulatory T cell proliferation. Immunity. 26:241–255. 2007. View Article : Google Scholar : PubMed/NCBI

42 

Li M, Feurino LW, Li F, Wang H, Zhai Q, Fisher WE, Chen C and Yao Q: Thymosinalpha1 stimulates cell proliferation by activating ERK1/2, JNK, and increasing cytokine secretion in human pancreatic cancer cells. Cancer Lett. 248:58–67. 2007. View Article : Google Scholar : PubMed/NCBI

43 

Li H, Cai X, Fan X, Moquin B, Stoicov C and Houghton J: Fas Ag-FasL coupling leads to ERK1/2-mediated proliferation of gastric mucosal cells. Am J Physiol Gastrointest Liver Physiol. 294:G263–G275. 2008. View Article : Google Scholar : PubMed/NCBI

44 

He Z, Jiang J, Kokkinaki M, Golestaneh N, Hofmann MC and Dym M: Gdnf upregulates c-Fos transcription via the Ras/Erk1/2 pathway to promote mouse spermatogonial stem cell proliferation. Stem Cells. 26:266–278. 2008. View Article : Google Scholar : PubMed/NCBI

45 

Mancinelli R, Onori P, Gaudio E, DeMorrow S, Franchitto A, Francis H, Glaser S, Carpino G, Venter J, Alvaro D, et al: Follicle-stimulating hormone increases cholangiocyte proliferation by an autocrine mechanism via cAMP-dependent phosphorylation of ERK1/2 and Elk-1. Am J Physiol Gastrointest Liver Physiol. 297:G11–G26. 2009. View Article : Google Scholar : PubMed/NCBI

46 

Sirianni R, Chimento A, De Luca A, Casaburi I, Rizza P, Onofrio A, Iacopetta D, Puoci F, Andò S, Maggiolini M and Pezzi V: Oleuropein and hydroxytyrosol inhibit MCF-7 breast cancer cell proliferation interfering with ERK1/2 activation. Mol Nutr. 54:833–840. 2010. View Article : Google Scholar

47 

Yang Y and Han C: GDNF stimulates the proliferation of cultured mouse immature Sertoli cells via its receptor subunit NCAM and ERK1/2 signaling pathway. BMC Cell≠≠≠ Bio≠≠.

48 

Lee JG and Kay EP: PI 3-kinase/Rac1 and ERK1/2 regulate FGF-2-mediated cell proliferation through phosphorylation of p27 at Ser10 by KIS and at Thr187 by Cdc25A/Cdk2. Invest Ophthalmol Vis Sci. 52:417–426. 2011. View Article : Google Scholar : PubMed/NCBI

49 

Gao M, Zhan YQ, Yu M, Ge CH, Li CY, Zhang JH, Wang XH, Ge ZQ and Yang XM: Hepassocin activates the EGFR/ERK cascade and induces proliferation of L02 cells through the Src-dependent pathway. Cell Signal. 26:2161–2166. 2014. View Article : Google Scholar : PubMed/NCBI

50 

Tocharus C, Puriboriboon Y, Junmanee T, Tocharus J, Ekthuwapranee K and Govitrapong P: Melatonin enhances adult rat hippocampal progenitor cell proliferation via ERK signaling pathway through melatonin receptor. Neuroscience. 275:314–321. 2014. View Article : Google Scholar : PubMed/NCBI

51 

Wu Z, Uchi H, Morino-Koga S, Shi W and Furue M: Resveratrol inhibition of human keratinocyte proliferation via SIRT1/ARNT/ERK dependent downregulation of aquaporin 3. J Dermatol Sci. 75:16–23. 2014. View Article : Google Scholar : PubMed/NCBI

52 

Liu H, Wu Y, Zhu S, Liang W, Wang Z, Wang Y, Lv T, Yao Y, Yuan D and Song Y: PTP1B promotes cell proliferation and metastasis through activating src and ERK1/2 in non-small cell lung cancer. Cancer Lett. 359:218–225. 2015. View Article : Google Scholar : PubMed/NCBI

53 

Wang J, He C, Zhou T, Huang Z, Zhou L and Liu X: NGF increases VEGF expression and promotes cell proliferation via ERK1/2 and AKT signaling in Müller cells. Mol Vis. 22:254–263. 2016.PubMed/NCBI

54 

Kim SH, Pei QM, Jiang P, Yang M, Qian XJ and Liu JB: Effect of active vitamin D3 on VEGF-induced ADAM33 expression and proliferation in human airway smooth muscle cells: Implications for asthma treatment. Respir Res. 18:72017. View Article : Google Scholar : PubMed/NCBI

55 

Graves LM, Guy HI, Kozlowski P, Huang M, Lazarowski E, Pope RM, Collins MA, Dahlstrand EN, Earp HS III and Evans DR: Regulation of carbamoyl phosphate synthetase by MAP kinase. Nature. 403:328–332. 2000. View Article : Google Scholar : PubMed/NCBI

56 

Stefanovsky V, Langlois F, Gagnon-Kugler T, Rothblum LI and Moss T: Growth factor signaling regulates elongation of RNA polymerase I transcription in mammals via UBF phosphorylation and r-chromatin remodeling. Mol Cell. 21:629–639. 2006. View Article : Google Scholar : PubMed/NCBI

57 

Mendoza MC, Er EE and Blenis J: The Ras-ERK and PI3K-mTOR pathways: Cross-talk and compensation. Trends Biochem Sci. 36:320–328. 2011. View Article : Google Scholar : PubMed/NCBI

58 

Wang Y, Zhu L, Kuokkanen S and Pollard JW: Activation of protein synthesis in mouse uterine epithelial cells by estradiol-17β is mediated by a PKC-ERK1/2-mTOR signaling pathway. Proc Natl Acad Sci USA. 112:E1382–E1391. 2015. View Article : Google Scholar : PubMed/NCBI

59 

Hoang B, Benavides A, Shi Y, Yang Y, Frost P, Gera J and Lichtenstein A: The PP242 mammalian target of rapamycin (mTOR) inhibitor activates extracellular signal-regulated kinase (ERK) in multiple myeloma cells via a target of rapamycin complex 1 (TORC1)/eukaryotic translation initiation factor 4E (eIF-4E)/RAF pathway and activation is a mechanism of resistance. J Biol Chem. 287:21796–21805. 2012. View Article : Google Scholar : PubMed/NCBI

60 

Ma L, Chen Z, Erdjument-Bromage H, Tempst P and Pandolfi PP: Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis. Cell. 121:179–193. 2005. View Article : Google Scholar : PubMed/NCBI

61 

Chambard JC, Lefloch R, Pouysségur J and Lenormand P: ERK implication in cell cycle regulation. Biochim Biophys Acta. 1773:1299–1310. 2007. View Article : Google Scholar : PubMed/NCBI

62 

Lavoie JN, L'Allemain G, Brunet A, Muller R and Pouysségur J: Cyclin D1 expression is regulated positively by the p42/p44MAPK and negatively by the p38/HOGMAPK pathway. J Biol Chem. 271:20608–20616. 1996. View Article : Google Scholar : PubMed/NCBI

63 

Seth A, Alvarez E, Gupta S and Davis RJ: A phosphorylation site located in the NH2-terminal domain of c-Myc increases transactivation of gene expression. J Biol Chem. 266:23521–23524. 1991.PubMed/NCBI

64 

Daksis JI, Lu RY, Facchini LM, Marhin WW and Penn LJ: Myc induces cyclin D1 expression in the absence of de novo protein synthesis and links mitogen-stimulated signal transduction to the cell cycle. Oncogene. 9:3635–3645. 1994.PubMed/NCBI

65 

Walsh S, Margolis SS and Kornbluth S: Phosphorylation of the cyclin B1 cytoplasmic retention sequence by mitogen-activated protein kinase and Plx. Mol Cancer Res. 1:280–289. 2003.PubMed/NCBI

66 

Palmer A, Gavin AC and Nebreda AR: A link between MAP kinase and p34(cdc2)/cyclin B during oocyte maturation: p90(rsk) phosphorylates and inactivates the p34(cdc2) inhibitory kinase Myt1. EMBO J. 17:5037–5047. 1998. View Article : Google Scholar : PubMed/NCBI

67 

Shaw PH: The role of p53 in cell cycle regulation. Pathol Res Pract. 192:669–675. 1996. View Article : Google Scholar : PubMed/NCBI

68 

Wesierska-Gadek J, Wojciechowski J, Ranftler C and Schmid G: Role of p53 tumor suppressor in ageing: Regulation of transient cell cycle arrest and terminal senescence. J Physiol Pharmacol. 56:15–28. 2005.PubMed/NCBI

69 

Lee SY, Choi HC, Choe YJ, Shin SJ, Lee SH and Kim HS: Nutlin-3 induces BCL2A1 expression by activating ELK1 through the mitochondrial p53-ROS-ERK1/2 pathway. Int J Oncol. 45:675–682. 2014. View Article : Google Scholar : PubMed/NCBI

70 

Murase S, Kim E, Lin L, Hoffman DA and McKay RD: Loss of signal transducer and activator of transcription 3 (STAT3) signaling during elevated activity causes vulnerability in hippocampal neurons. J Neurosci. 32:15511–15520. 2012. View Article : Google Scholar : PubMed/NCBI

71 

Carlos AR, Escandell JM, Kotsantis P, Suwaki N, Bouwman P, Badie S, Folio C, Benitez J, Gomez-Lopez G, Pisano DG, et al: ARF triggers senescence in Brca2-deficient cells by altering the spectrum of p53 transcriptional targets. Nat Commun. 4:26972013. View Article : Google Scholar : PubMed/NCBI

72 

Tang D, Wu D, Hirao A, Lahti JM, Liu L, Mazza B, Kidd VJ, Mak TW and Ingram AJ: ERK activation mediates cell cycle arrest and apoptosis after DNA damage independently of p53. J Biol Chem. 277:12710–12717. 2002. View Article : Google Scholar : PubMed/NCBI

73 

Ashcroft M and Vousden KH: Regulation of p53 stability. Oncogene. 18:7637–7643. 1999. View Article : Google Scholar : PubMed/NCBI

74 

Ling Q, Meng C, Chen Q and Xing D: Activated ERK/FOXM1 pathway by low-power laser irradiation inhibits UVB-induced senescence through down-regulating p21 expression. J Cell Physiol. 229:108–116. 2014.PubMed/NCBI

75 

Rasola A, Sciacovelli M, Pantic B and Bernardi P: Signal transduction to the permeability transition pore. FEBS Lett. 584:1989–1996. 2010. View Article : Google Scholar : PubMed/NCBI

76 

Kashatus JA, Nascimento A, Myers LJ, Sher A, Byrne FL, Hoehn KL, Counter CM and Kashatus DF: Erk2 phosphorylation of Drp1 promotes mitochondrial fission and MAPK-driven tumor growth. Mol Cell. 57:537–551. 2015. View Article : Google Scholar : PubMed/NCBI

77 

Zhu J, Woods D, McMahon M and Bishop JM: Senescence of human fibroblasts induced by oncogenic Raf. Genes Dev. 12:2997–3007. 1998. View Article : Google Scholar : PubMed/NCBI

78 

Cammarano MS, Nekrasova T, Noel B and Minden A: Pak4 induces premature senescence via a pathway requiring p16INK4/p19ARF and mitogen-activated protein kinase signaling. Mol Cell Biol. 25:9532–9542. 2005. View Article : Google Scholar : PubMed/NCBI

79 

Kim-Kaneyama, Nose K and Shibanuma M: Significance of nuclear relocalization of ERK1/2 in reactivation of c-fos transcription and DNA synthesis in senescent fibroblasts. J Biol Chem. 275:20685–20692. 2000. View Article : Google Scholar : PubMed/NCBI

80 

Lim IK, Won Hong K, Kwak IH, Yoon G and Park SC: Cytoplasmic retention of p-Erk1/2 and nuclear accumulation of actin proteins during cellular senescence in human diploid fibroblasts. Mech Ageing Dev. 119:113–130. 2000. View Article : Google Scholar : PubMed/NCBI

81 

Chaturvedi V, Cesnjaj M, Bacon P, Panella J, Choubey D, Diaz MO and Nickoloff BJ: Role of INK4a/Arf locus-encoded senescent checkpoints activated in normal and psoriatic keratinocytes. Am J Pathol. 162:161–170. 2003. View Article : Google Scholar : PubMed/NCBI

82 

Kim HS, Song MC, Kwak IH, Park TJ and Lim IK: Constitutive induction of p-Erk1/2 accompanied by reduced activities of protein phosphatases 1 and 2A and MKP3 due to reactive oxygen species during cellular senescence. J Biol Chem. 278:37497–37510. 2003. View Article : Google Scholar : PubMed/NCBI

83 

Todd DE, Densham RM, Molton SA, Balmanno K, Newson C, Weston CR, Garner AP, Scott L and Cook SJ: ERK1/2 and p38 cooperate to induce a p21CIP1-dependent G1 cell cycle arrest. Oncogene. 23:3284–3295. 2004. View Article : Google Scholar : PubMed/NCBI

84 

Klein LE, Freeze BS, Smith AB III and Horwitz SB: The microtubule stabilizing agent discodermolide is a potent inducer of accelerated cell senescence. Cell Cycle. 4:501–507. 2005. View Article : Google Scholar : PubMed/NCBI

85 

Albrecht DS, Clubbs EA, Ferruzzi M and Bomser JA: Epigallocatechin-3-gallate (EGCG) inhibits PC-3 prostate cancer cell proliferation via MEK-independent ERK1/2 activation. Chem Biol Interact. 171:89–95. 2008. View Article : Google Scholar : PubMed/NCBI

86 

Deschênes-Simard X, Gaumont-Leclerc MF, Bourdeau V, Lessard F, Moiseeva O, Forest V, Igelmann S, Mallette FA, Saba-El-Leil MK, Meloche S, et al: Tumor suppressor activity of the ERK/MAPK pathway by promoting selective protein degradation. Genes Dev. 27:900–915. 2013. View Article : Google Scholar : PubMed/NCBI

87 

Zhu B, Ferry CH, Blazanin N, Bility MT, Khozoie C, Kang BH, Glick AB, Gonzalez FJ and Peters JM: PPARβ/δ promotes HRAS-induced senescence and tumor suppression by potentiating p-ERK and repressing p-AKT signaling. Oncogene. 33:5348–5359. 2014. View Article : Google Scholar : PubMed/NCBI

88 

Wang Z, Liu Y, Takahashi M, Van Hook K, Kampa-Schittenhelm KM, Sheppard BC, Sears RC, Stork PJ and Lopez CD: N terminus of ASPP2 binds to Ras and enhances Ras/Raf/MEK/ERK activation to promote oncogene-induced senescence. Proc Natl Acad Sci USA. 110:312–317. 2013. View Article : Google Scholar : PubMed/NCBI

89 

El Bezawy R, De Cesare M, Pennati M, Deraco M, Gandellini P, Zuco V and Zaffaroni N: Antitumor activity of miR-34a in peritoneal mesothelioma relies on c-MET and AXL inhibition: Persistent activation of ERK and AKT signaling as a possible cytoprotective mechanism. J Hematol Oncol. 10:192017. View Article : Google Scholar : PubMed/NCBI

90 

del Nogal M, Troyano N, Calleros L, Griera M, Rodriguez-Puyol M, Rodriguez-Puyol D and Ruiz-Torres MP: Hyperosmolarity induced by high glucose promotes senescence in human glomerular mesangial cells. Int J Biochem Cell Biol. 54:98–110. 2014. View Article : Google Scholar : PubMed/NCBI

91 

Lake D, Correa SA and Müller J: Negative feedback regulation of the ERK1/2 MAPK pathway. Cell Mol Life Sci. 73:4397–4413. 2016. View Article : Google Scholar : PubMed/NCBI

92 

Northwood IC, Gonzalez FA, Wartmann M, Raden DL and Davis RJ: Isolation and characterization of two growth factor-stimulated protein kinases that phosphorylate the epidermal growth factor receptor at threonine 669. J Biol Chem. 266:15266–15276. 1991.PubMed/NCBI

93 

Sato K, Shin MS, Sakimura A, Zhou Y, Tanaka T, Kawanishi M, Kawasaki Y, Yokoyama S, Koizumi K, Saiki I and Sakurai H: Inverse correlation between Thr-669 and constitutive tyrosine phosphorylation in the asymmetric epidermal growth factor receptor dimer conformation. Cancer Sci. 104:1315–1322. 2013. View Article : Google Scholar : PubMed/NCBI

94 

Prahallad A, Sun C, Huang S, Di Nicolantonio F, Salazar R, Zecchin D, Beijersbergen RL, Bardelli A and Bernards R: Unresponsiveness of colon cancer to BRAF(V600E) inhibition through feedback activation of EGFR. Nature. 483:100–103. 2012. View Article : Google Scholar : PubMed/NCBI

95 

Zakrzewska M, Haugsten EM, Nadratowska-Wesolowska B, Oppelt A, Hausott B, Jin Y, Otlewski J, Wesche J and Wiedlocha A: ERK-mediated phosphorylation of fibroblast growth factor receptor 1 on Ser777 inhibits signaling. Sci Signal. 6:ra112013. View Article : Google Scholar : PubMed/NCBI

96 

Kamioka Y, Yasuda S, Fujita Y, Aoki K and Matsuda M: Multiple decisive phosphorylation sites for the negative feedback regulation of SOS1 via ERK. J Biol Chem. 285:33540–33548. 2010. View Article : Google Scholar : PubMed/NCBI

97 

Lax I, Wong A, Lamothe B, Lee A, Frost A, Hawes J and Schlessinger J: The docking protein FRS2alpha controls a MAP kinase-mediated negative feedback mechanism for signaling by FGF receptors. Mol Cell. 10:709–719. 2002. View Article : Google Scholar : PubMed/NCBI

98 

Wu YJ, Chen ZJ and Ullrich A: EGFR and FGFR signaling through FRS2 is subject to negative feedback control by ERK1/2. Biol Chem. 384:1215–1226. 2003. View Article : Google Scholar : PubMed/NCBI

99 

Wartmann M, Hofer P, Turowski P, Saltiel AR and Hynes NE: Negative modulation of membrane localization of the Raf-1 protein kinase by hyperphosphorylation. J Biol Chem. 272:3915–3923. 1997. View Article : Google Scholar : PubMed/NCBI

100 

Weiss RH, Maga EA and Ramirez A: MEK inhibition augments Raf activity, but has variable effects on mitogenesis, in vascular smooth muscle cells. Am J Physiol. 274:C1521–C1529. 1998. View Article : Google Scholar : PubMed/NCBI

101 

Dougherty MK, Müller J, Ritt DA, Zhou M, Zhou XZ, Copeland TD, Conrads TP, Veenstra TD, Lu KP and Morrison DK: Regulation of Raf-1 by direct feedback phosphorylation. Mol Cell. 17:215–224. 2005. View Article : Google Scholar : PubMed/NCBI

102 

Hekman M, Fischer A, Wennogle LP, Wang YK, Campbell SL and Rapp UR: Novel C-Raf phosphorylation sites: Serine 296 and 301 participate in Raf regulation. FEBS Lett. 579:464–468. 2005. View Article : Google Scholar : PubMed/NCBI

103 

Balan V, Leicht DT, Zhu J, Balan K, Kaplun A, Singh-Gupta V, Qin J, Ruan H, Comb MJ and Tzivion G: Identification of novel in vivo Raf-1 phosphorylation sites mediating positive feedback Raf-1 regulation by extracellular signal-regulated kinase. Mol Biol Cell. 17:1141–1153. 2006. View Article : Google Scholar : PubMed/NCBI

104 

Brummer T, Naegele H, Reth M and Misawa Y: Identification of novel ERK-mediated feedback phosphorylation sites at the C-terminus of B-Raf. Oncogene. 22:8823–8834. 2003. View Article : Google Scholar : PubMed/NCBI

105 

Ritt DA, Monson DM, Specht SI and Morrison DK: Impact of feedback phosphorylation and Raf heterodimerization on normal and mutant B-Raf signaling. Mol Cell Biol. 30:806–819. 2010. View Article : Google Scholar : PubMed/NCBI

106 

Eblen ST, Slack-Davis JK, Tarcsafalvi A, Parsons JT, Weber MJ and Catling AD: Mitogen-activated protein kinase feedback phosphorylation regulates MEK1 complex formation and activation during cellular adhesion. Mol Cell Biol. 24:2308–2317. 2004. View Article : Google Scholar : PubMed/NCBI

107 

Rossomando AJ, Dent P, Sturgill TW and Marshak DR: Mitogen-activated protein kinase kinase 1 (MKK1) is negatively regulated by threonine phosphorylation. Mol Cell Biol. 14:1594–1602. 1994. View Article : Google Scholar : PubMed/NCBI

108 

Canal F, Palygin O, Pankratov Y, Corrêa SA and Müller J: Compartmentalization of the MAPK scaffold protein KSR1 modulates synaptic plasticity in hippocampal neurons. FASEB J. 25:2362–2372. 2011. View Article : Google Scholar : PubMed/NCBI

109 

McKay MM, Ritt DA and Morrison DK: Signaling dynamics of the KSR1 scaffold complex. Proc Natl Acad Sci USA. 106:11022–11027. 2009. View Article : Google Scholar : PubMed/NCBI

110 

Fey D, Croucher DR, Kolch W and Kholodenko BN: Crosstalk and signaling switches in mitogen-activated protein kinase cascades. Front Physiol. 3:3552012. View Article : Google Scholar : PubMed/NCBI

111 

Caunt CJ, Finch AR, Sedgley KR and McArdle CA: Seven-transmembrane receptor signalling and ERK compartmentalization. Trends Endocrinol Metab. 17:276–283. 2006. View Article : Google Scholar : PubMed/NCBI

112 

Owens DM and Keyse SM: Differential regulation of MAP kinase signalling by dual-specificity protein phosphatases. Oncogene. 26:3203–3213. 2007. View Article : Google Scholar : PubMed/NCBI

113 

Huang CY and Tan TH: DUSPs, to MAP kinases and beyond. Cell Biosci. 2:242012. View Article : Google Scholar : PubMed/NCBI

114 

Peti W and Page R: Molecular basis of MAP kinase regulation. Protein Sci. 22:1698–1710. 2013. View Article : Google Scholar : PubMed/NCBI

115 

Tanzola MB and Kersh GJ: The dual specificity phosphatase transcriptome of the murine thymus. Mol Immunol. 43:754–762. 2006. View Article : Google Scholar : PubMed/NCBI

116 

Hanafusa H, Torii S, Yasunaga T and Nishida E: Sprouty1 and Sprouty2 provide a control mechanism for the Ras/MAPK signalling pathway. Nat Cell Biol. 4:850–858. 2002. View Article : Google Scholar : PubMed/NCBI

117 

Lito P, Pratilas CA, Joseph EW, Tadi M, Halilovic E, Zubrowski M, Huang A, Wong WL, Callahan MK, Merghoub T, et al: Relief of profound feedback inhibition of mitogenic signaling by RAF inhibitors attenuates their activity in BRAFV600E melanomas. Cancer Cell. 22:668–682. 2012. View Article : Google Scholar : PubMed/NCBI

118 

Lito P, Rosen N and Solit DB: Tumor adaptation and resistance to RAF inhibitors. Nat Med. 19:1401–1409. 2013. View Article : Google Scholar : PubMed/NCBI

119 

Yusoff P, Lao DH, Ong SH, Wong ES, Lim J, Lo TL, Leong HF, Fong CW and Guy GR: Sprouty2 inhibits the Ras/MAP kinase pathway by inhibiting the activation of Raf. J Biol Chem. 277:3195–3201. 2002. View Article : Google Scholar : PubMed/NCBI

120 

Dent P: Crosstalk between ERK, AKT, and cell survival. Cancer Biol Ther. 15:245–246. 2014. View Article : Google Scholar : PubMed/NCBI

121 

Mace PD, Wallez Y, Egger MF, Dobaczewska MK, Robinson H, Pasquale EB and Riedl SJ: Structure of ERK2 bound to PEA-15 reveals a mechanism for rapid release of activated MAPK. Nat Commun. 4:16812013. View Article : Google Scholar : PubMed/NCBI

122 

Sinha D, Bannergee S, Schwartz JH, Lieberthal W and Levine JS: Inhibition of ligand-independent ERK1/2 activity in kidney proximal tubular cells deprived of soluble survival factors up-regulates Akt and prevents apoptosis. J Biol Chem. 279:10962–10972. 2004. View Article : Google Scholar : PubMed/NCBI

123 

Trencia A, Perfetti A, Cassese A, Vigliotta G, Miele C, Oriente F, Santopietro S, Giacco F, Condorelli G, Formisano P and Beguinot F: Protein kinase B/Akt binds and phosphorylates PED/PEA-15, stabilizing its antiapoptotic action. Mol Cell Biol. 23:4511–4521. 2003. View Article : Google Scholar : PubMed/NCBI

124 

Aksamitiene E, Kiyatkin A and Kholodenko BN: Cross-talk between mitogenic Ras/MAPK and survival PI3K/Akt pathways: A fine balance. Biochem Soc Trans. 40:139–146. 2012. View Article : Google Scholar : PubMed/NCBI

125 

Ussar S and Voss T: MEK1 and MEK2, different regulators of the G1/S transition. J Biol Chem. 279:43861–43869. 2004. View Article : Google Scholar : PubMed/NCBI

126 

Lee SJ, Lee SH, Yoon MH and Park BJ: A new p53 target gene, RKIP, is essential for DNA damage-induced cellular senescence and suppression of ERK activation. Neoplasia. 15:727–737. 2013. View Article : Google Scholar : PubMed/NCBI

127 

Li M, Zhou JY, Ge Y, Matherly LH and Wu GS: The phosphatase MKP1 is a transcriptional target of p53 involved in cell cycle regulation. J Biol Chem. 278:41059–41068. 2003. View Article : Google Scholar : PubMed/NCBI

128 

Shen WH, Wang J, Wu J, Zhurkin VB and Yin Y: Mitogen-activated protein kinase phosphatase 2: A novel transcription target of p53 in apoptosis. Cancer Res. 66:6033–6039. 2006. View Article : Google Scholar : PubMed/NCBI

129 

Ueda K, Arakawa H and Nakamura Y: Dual-specificity phosphatase 5 (DUSP5) as a direct transcriptional target of tumor suppressor p53. Oncogene. 22:5586–5591. 2003. View Article : Google Scholar : PubMed/NCBI

130 

El Hasasna H, Athamneh K, Al Samri H, Karuvantevida N, Al Dhaheri Y, Hisaindee S, Ramadan G, Al Tamimi N, AbuQamar S, Eid A and Iratni R: Rhus coriaria induces senescence and autophagic cell death in breast cancer cells through a mechanism involving p38 and ERK1/2 activation. Sci Rep. 5:130132015. View Article : Google Scholar : PubMed/NCBI

131 

Wajapeyee N, Serra RW, Zhu X, Mahalingam M and Green MR: Oncogenic BRAF induces senescence and apoptosis through pathways mediated by the secreted protein IGFBP7. Cell. 132:363–374. 2008. View Article : Google Scholar : PubMed/NCBI

132 

Ichimura A, Ruike Y, Terasawa K, Shimizu K and Tsujimoto G: MicroRNA-34a inhibits cell proliferation by repressing mitogen-activated protein kinase kinase 1 during megakaryocytic differentiation of K562 cells. Mol Pharmacol. 77:1016–1024. 2010. View Article : Google Scholar : PubMed/NCBI

133 

Sayed D, Rane S, Lypowy J, He M, Chen IY, Vashistha H, Yan L, Malhotra A, Vatner D and Abdellatif M: MicroRNA-21 targets Sprouty2 and promotes cellular outgrowths. Mol Biol Cell. 19:3272–3282. 2008. View Article : Google Scholar : PubMed/NCBI

134 

Deschênes-Simard X, Kottakis F, Lessard F, Saint-Germain E, Bourdeau VBardeesy N and Ferbeyre G: Tumor suppressor activity of the ERK/MAPK signaling: Inhibition of cell reprogramming by degradation of specific proteins. Cancer Res. 74:38952014. View Article : Google Scholar

135 

Plotnikov A, Flores K, Maik-Rachline G, Zehorai E, Kapri-Pardes E, Berti DA, Hanoch T, Besser MJ and Seger R: The nuclear translocation of ERK1/2 as an anticancer target. Nat Commun. 6:66852015. View Article : Google Scholar : PubMed/NCBI

136 

Schevzov G, Kee AJ, Wang B, Sequeira VB, Hook J, Coombes JD, Lucas CA, Stehn JR, Musgrove EA, Cretu A, et al: Regulation of cell proliferation by ERK and signal-dependent nuclear translocation of ERK is dependent on Tm5NM1-containing actin filaments. Mol Biol Cell. 26:2475–2490. 2015. View Article : Google Scholar : PubMed/NCBI

137 

Wainstein E and Seger R: The dynamic subcellular localization of ERK: Mechanisms of translocation and role in various organelles. Curr Opin Cell Biol. 39:15–20. 2016. View Article : Google Scholar : PubMed/NCBI

138 

Callejas-Valera JL, Guinea-Viniegra J, Ramirez-Castillejo C, Recio JA, Galan-Moya E, Martinez N, Rojas JM, Ramón y Cajal S and Sánchez-Prieto R: E1a gene expression blocks the ERK1/2 signaling pathway by promoting nuclear localization and MKP up-regulation: Implication in v-H-Ras-induced senescence. J Biol Chem. 283:13450–13458. 2008. View Article : Google Scholar : PubMed/NCBI

139 

Gaumont-Leclerc MF, Mukhopadhyay UK, Goumard S and Ferbeyre G: PEA-15 is inhibited by adenovirus E1A and plays a role in ERK nuclear export and Ras-induced senescence. J Biol Chem. 279:46802–46809. 2004. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

February-2019
Volume 19 Issue 2

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zou J, Lei T, Guo P, Yu J, Xu Q, Luo Y, Ke R and Huang D: Mechanisms shaping the role of ERK1/2 in cellular senescence (Review). Mol Med Rep 19: 759-770, 2019
APA
Zou, J., Lei, T., Guo, P., Yu, J., Xu, Q., Luo, Y. ... Huang, D. (2019). Mechanisms shaping the role of ERK1/2 in cellular senescence (Review). Molecular Medicine Reports, 19, 759-770. https://doi.org/10.3892/mmr.2018.9712
MLA
Zou, J., Lei, T., Guo, P., Yu, J., Xu, Q., Luo, Y., Ke, R., Huang, D."Mechanisms shaping the role of ERK1/2 in cellular senescence (Review)". Molecular Medicine Reports 19.2 (2019): 759-770.
Chicago
Zou, J., Lei, T., Guo, P., Yu, J., Xu, Q., Luo, Y., Ke, R., Huang, D."Mechanisms shaping the role of ERK1/2 in cellular senescence (Review)". Molecular Medicine Reports 19, no. 2 (2019): 759-770. https://doi.org/10.3892/mmr.2018.9712