Journal list menu

Volume 277, Issue 1 p. 2-21
MINIREVIEW
Free Access

ERK and cell death: Mechanisms of ERK-induced cell death – apoptosis, autophagy and senescence

Sebastien Cagnol

Sebastien Cagnol

Department of Anatomy and Cellular Biology, Faculty of Medicine and Health Sciences, University of Sherbrooke, Canada

Search for more papers by this author
Jean-Claude Chambard

Jean-Claude Chambard

Institute of Developmental Biology and Cancer, University of Nice, France

Search for more papers by this author
First published: 15 December 2009
Citations: 963
S. Cagnol, Department of Anatomy and Cellular Biology, Faculty of Medicine and Health Sciences, University of Sherbrooke, Sherbrooke, Quebec, Canada
Fax: +1 819 564 5320
Tel: +1 819 820 6868 ext. 15715
E-mail: [email protected]

Abstract

The Ras/Raf/extracellular signal-regulated kinase (ERK) signaling pathway plays a crucial role in almost all cell functions and therefore requires exquisite control of its spatiotemporal activity. Depending on the cell type and stimulus, ERK activity will mediate different antiproliferative events, such as apoptosis, autophagy and senescence in vitro and in vivo. ERK activity can promote either intrinsic or extrinsic apoptotic pathways by induction of mitochondrial cytochrome c release or caspase-8 activation, permanent cell cycle arrest or autophagic vacuolization. These unusual effects require sustained ERK activity in specific subcellular compartments and could depend on the presence of reactive oxygen species. We will summarize the mechanisms involved in Ras/Raf/ERK antiproliferative functions.

Abbreviations

  • ATA
  • aurintricarboxylic acid
  • cPLA2
  • cytosolic phospholipase A2
  • DAPK
  • death-associated protein kinase
  • DUSP
  • dual-specificity phosphatase
  • EGF
  • epidermal growth factor
  • ERK
  • extracellular signal-regulated kinase
  • FADD
  • Fas-associated death domain
  • GAIP
  • G-interacting protein
  • IGF
  • insulin-like growth factor
  • JNK
  • c-JunNH2-terminal kinase
  • MAPK
  • mitogen-activated protein kinase
  • MEK1/2
  • mitogen protein kinase kinase 1 and 2 (also known as MAP2K1 and MAP2K2, respectively)
  • MEKCA
  • constitutively activated forms of MEK
  • MKP
  • mitogen-activated protein kinase phosphatase
  • MOS
  • v-mos Moloney murine sarcoma viral oncogene homolog
  • PARP
  • poly(ADP-ribose) polymerase
  • ROS
  • reactive oxygen species
  • TNF
  • tumor necrosis factor
  • TRAIL
  • tumor necrosis factor-related apoptosis-inducing ligand
  • Ras/Raf/ERK, the pathway

    ERK2/ERK1 (also known as p42/p44MAPK, respectively, and officially named MAPK 1 and 3) are two isoforms of extracellular signal-regulated kinase (ERK) that belong to the family of mitogen-activated protein kinases (MAPKs), which include ERK5, the c-JunNH2-terminal kinases (JNK1/2/3) and the p38 MAP kinases (p38 α,β,δ,γ). These enzymes are activated through a sequential phosphorylation cascade that amplifies and transduces signals from the cell membrane to the nucleus. Upon receptor activation, membrane-bound GTP-loaded Ras recruits one of the Raf kinases, A-Raf, B-Raf and C-Raf (or Raf1), into a complex where it becomes activated. Then, Raf phosphorylates two serine residues on the kinase mitogen protein kinase kinase 1 and 2 (MEK1/2; also known as MAP2K1 and MAP2K2, respectively), which in turn activate ERK1/2 by tandem phosphorylation of threonine and tyrosine residues on the dual-specificity motif (T-E-Y). Finally, active ERKs regulate by phosphorylation many cytoplasmic and nuclear targets that perform important biological functions [1].

    Depending on the duration, the magnitude and its subcellular localization, ERK activation controls various cell responses, such as proliferation, migration, differentiation and death [2]. Protein phosphatases play an important role as negative regulators by controlling the Ras/Raf/ERK signaling pathway at different levels. Phosphotyrosine phosphatases target the tyrosine kinase receptors, whereas phosphoserine/phosphothreonine phosphatases target the adapter protein Shc and MEK1/2. Dual-specificity phosphatases [DUSP; also called MAPK phosphatases (MKP)], are able to dephosphorylate both of the threonine and tyrosine residues within the activation loop of MAPK. Specific DUSPs tightly regulate subcellular ERK activity. DUSP1/MKP-1, DUSP2/PAC-1, DUSP4/MKP-2 and DUSP5 are mainly nuclear, whereas DUSP6/MKP-3, DUSP7/MKP-X and DUSP9/MKP-4 are cytoplasmic. Moreover, the expression of DUSP1, -2, -4 and -6 is increased following ERK activation [3,4], taking part in a negative feedback loop aimed at terminating Ras/Raf/ERK signaling pathway stimulation.

    The Ras/Raf/ERK pathway is frequently deregulated in tumors as a result of activating mutations in Ras or B-Raf, observed particularly in malignant melanoma, pancreas intestine and thyroid tumors (cosmic database: http://www.sanger.ac.uk/genetics/CGP/cosmic). Many studies associate its oncogenic potential to increased cell survival, mainly by promoting the activity of antiapoptotic proteins, such as Bcl-2, Bcl-XL, Mcl-1, IAP, and repressing proapoptotic proteins, such as Bad and Bim [5].

    Paradoxically, a growing number of studies also suggest that in certain conditions, aberrant ERK activation can promote cell death. This review will summarize the different cellular models in which the Ras/Raf/ERK pathway plays an antiproliferative role. The specific pro-death function of ERK activity in neurons [6] and lymphocytes [7] and its role in cadmium toxicity [8] will also be discussed in this minireview series.

    Ras/Raf/ERK pathway induces apoptosis

    Programmed cell death by apoptosis is a cell-autonomous mechanism that relies on pathway-controlled activation of caspases and nucleases leading to the death of the injured cells without affecting neighboring cells. The intrinsic pathway of apoptosis regulates the activity of the Bcl-2 family proteins that control the integrity of the mitochondrial membrane. The release of proapoptotic factors from the mitochondria, such as cytochrome c, into the cytoplasm promotes the activation of initiator caspase-9, which in turn activates effector caspases such as caspase-3 or -7. The extrinsic pathway relies on the activation of death receptors from the tumor necrosis factor (TNF) receptor family that promote the recruitment and activation of initiator caspase-8 via adaptor proteins such as Fas-associated death domain (FADD) or TNFRSF1A-associated via death domain (TRADD). Strong caspase-8 activity may directly activate effector caspases; it may also require signal amplification through induction of the intrinsic pathway via cleavage of the Bcl-2 family protein Bid [9].

    Early reports of a proapoptotic function of the Ras/Raf/ERK pathway appeared in 1996. Depletion of Raf by the benzoquinone ansamycin geldanamycin was shown to protect MCF-7 cells from apoptosis induced by the antitumor compound taxol [10], whereas MEK antisense cDNA expression prevented bufalin-induced apoptosis in U937 leukemic cells [11]. A growing number of studies using MEK inhibitors (PD98059, U0126) and expression of dominant negative or constitutively active forms of Ras, Raf, MEK or ERK have confirmed the implication of the Ras/Raf/ERK pathway in the induction of apoptosis (see Table 1 for details).

    Table 1. Models of ERK-mediated apoptosis and autophagy. LDH, lactate dehydrogenase; MDC, monodansylcadaverine; MTT, 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-tetrazolium bromide; ND, not determined; TGHQ, 2,3,5-tris-(glutathione-S-yl)hydroquinone; TPA, 12-O-tetradecanoylphorbol-13-acetate.
    In vivo/cellular model Stimuli inducing cell death Duration of ERK activation promoting cell death Characteristics of cell death Evidence implicating MEK-ERK in cell death Reference
    Transformed mouse fibroblast Etoposide 24 h Caspase-3 PD98059 [12]
    NIH 3T3 Etoposide
    UV
    Doxorubicin
    24 h DNA degradation PD98059 [13]
    Human keratinocytes HaCaT Etoposide 24 h DNA condensation Caspase-3
    PARP cleavage
    PD98059
    DN ERK
    [14]
    Human hepatocellular carcinoma HepG2 and Huh-7 Doxorubicin ND PARP cleavage PD98059 [16]
    Human promonocytic leukemia TPA
    ArsenicCadmium
    Doxorubicin
    24 h DNA fragmentation
    DNA condensation
    U0126
    PD98059
    [17]
    Human breast adenocarcinoma MCF-7 Doxorubicin 12 h Cell viability U0126 [18]
    NIH 3T3, human immortalized keratinocytes HaCaT Doxorubicin 24 h DNA fragmentation
    PARP cleavage
    PD98059
    DN ERK
    [19]
    Rat immortalized cardiomyocytes H9c2 Doxorubicin 48 h nuclear DNA fragmentation
    Caspase-9, -3
    PARP cleavage
    U0126 [20]
    NIH 3T3 γ irradiation 48 h Membrane integrity
    Annexin V
    PD98059
    DN ERK
    [21]
    Human cervix adenocarcinoma HeLa
    Human lung carcinoma A549
    Cisplatin 20 h DNA condensation
    Caspase-3
    PARP cleavage
    Cytochrome c release
    U0126
    PD98059
    [22]
    Human ovarian adenocarcinoma A2780 Cisplatin ND DNA fragmentation PD98059 [23]
    Human osteosarcoma Saos-2 Cisplatin 24 h Cell viability
    DNA fragmentation
    U0126
    PD98059
    [24]
    Rabbit primary renal proximal tubular cells Cisplatin 24 h DNA condensation
    Caspase-3
    U0126
    PD98059
    [25]
    Mouse immortalized proximal tubule cell line (TKPTS) Cisplatin 72 h Caspase-3 U0126 [26]
    Human carcinoma NCCIT and NTERA Cisplatin 24 h DNA condensation
    Caspase-8, -9, -3
    U0126
    PD98059
    [27]
    In vivo mouse kidney Cisplatin injection 72 h DNA fragmentation
    Caspase-8, -3
    U0126 [28]
    Opossum immortalized kidney cells OK cells Cisplatin 48 h DNA degradation
    Caspase-3
    U0126
    PD98059
    DN MEK
    [29]
    Human cervix adenocarcinoma HeLa Cisplatin 18 h Caspase-9
    PARP cleavage
    U0126 [30]
    Human papillary thyroid carcinoma BHP 2–7 and BHP 18–21 Resveratrol 24 h nuclear DNA degradation PD98059 [33]
    Human prostate adenocarcinoma PC3 Phenethyl isothiocyanate 24 h Annexin V
    ROS production
    PD98059 [35]
    Human melanoma C8161,WM164 Mel Juso Betulinic acid 96 h DNA fragmentation
    DNA condensation
    PARP cleavage
    U0126 [36]
    Human cervix adenocarcinoma HeLa Apigenin 8 h Cell viability U0126
    PD98059
    [37]
    Human melanoma A375-S2 Oridonin 12 h DNA fragmentation
    Cytochrome c release
    PD98059 [38]
    Human glioblastoma T98G and U87MG Miltefosine 12 h Cell viability U0126 [39]
    Human cervix adenocarcinoma HeLa Shikonin 12 h Caspase-8, -3 PD98059 [40]
    Human breast adenocarcinoma MCF-7 Taxol 24 h DNA fragmentation PD98059 [41]
    Normal human embryonic kidney HEK TRAIL Constitutive DNA fragmentation
    Caspase-8
    U0126
    PD98059
    [42]
    Human primary fibroblast BJ TRAIL ND cell viability U0126 [43]
    Human colorectal HT29 cells TRAIL 5 h Membrane integrity
    Nuclear condensation
    PARP cleavage
    PD98059 [44]
    Human prostate cancer LNCaP TRAIL 4 h Annexin V
    Caspase-3
    U0126 [45]
    Human prostate tumor DU-145 Quercetin
    TRAIL
    24 h Membrane integrity PD98059 [34]
    Human neuroblastoma SHEP-1 TRAIL
    H2O2
    ND Membrane integrity PD98059 [46]
    Huaman neuroblastoma SHEP FasL 30 min DNA condensation DN MEK1 [50]
    Rat primary Sertoli cells
    Human acute T leukemia Jurkat
    Fas (CH11) 5 min Membrane integrity
    DNA fragmentation
    PD98059 [51]
    Diffuse large B-cell lymphoma CD40 ligation 3 h Membrane integrity
    DNA fragmentation
    U0126
    PD98059
    [52]
    Primary human cholangiocytes CD40 ligation 24 h DNA condensation
    Caspase-3 activity
    PD98059 [53]
    Pig renal tubular epithelial cells LLC·PK1 Zinc 24 h nuclear Cell viabilityROS production U0126 [54]
    In vivo/isolated rat renal cortical slices ZnCl2 injection 90 min nuclear Cell viability U0126 [55]
    Pig renal tubular epithelial cells LLC·PK1 TGHQ 5 h Cell viability PD98059 [56]
    Rabbit primary renal proximal tubular cells tert-butylhydroperoxide 8 h Annexin V U0126
    PD98059
    [58]
    Human primary retinal pigmented epithelial ARPE19 cells tert-butylhydroperoxide 6 h Cell viability
    Caspase-9
    DNA fragmentation
    U0126 [59]
    Murine transformed lung epithelial MLE12 Hyperoxia 4 h Caspase-9, -3
    Cytochrome c release
    PD98059 [57]
    In vivo mouse lung Hyperoxia 72 h DNA fragmentation
    Caspase-3
    PD98059 [57]
    Primary rat pulmonary myofibroblasts ONOO 30 min Cell viability PD98059 [60]
    Immortalized mouse fibroblast L929 H2O2 3 h DNA fragmentation
    Annexin V
    PD98059
    DN ERK
    [61]
    Mouse immortalized osteoblast H2O2 Biphasic 12 h Cell viability
    Membrane integrity
    PD98059 [62]
    Rabbit primary renal proximal tubular cells H2O2 2 h constitutive DNA condensation
    Caspase-3
    U0126
    PD98059
    MEKCA
    [63]
    Rabbit primary renal proximal tubular cells H2O2 2 h Membrane integrity
    Annexin V
    U0126
    PD98059
    MEKCA
    DN MEK
    [64]
    Human transformed bronchial epithelial cell line BEAS-2B NO donor 24 h Cell viability PD98059 [65]
    Human melanoma NO donor ND PARP cleavage U0126 [67]
    HEK293 Cadmium Biphasic 96 h Caspase-8, -3
    PARP
    U0126 [68]
    Human hepatocellular carcinoma HepG2 Benzo[a]pyrene 48 h Cell viability PD98059 [70]
    Rat primary pleural mesothelial cells Crocidotite
    Asbestos
    48 h DNA condensation
    ROS production
    PD98059 [71]
    In vivo Caenorhabditis elegans Arsenic ND DNA fragmentation mek-2 (n1989)
    mpk-1 (ku1)
    [72]
    Murin bone marrow-derived primary osteoclasts and murine long bone-derived osteocytes MLO-Y4 Estradiol 24 h Membrane integrity U0126 [73]
    Human breast adenocarcinoma MCF-7 Tamoxifen 1 h Membrane integrity PD98059 [74]
    Human myeloma cell line U266-1984 and RHEK-1 Interferon-α 16 h Annexin V
    Caspase-3
    U0126 [75]
    Isolated rat renal cortical slices Cephaloridine 90 min nuclear Cell viability
    ROS production
    U0126
    PD98059
    [76]
    Immortalized rabbit lens epithelial cells N/N1003A Calcimycin 10 h Membrane integrity
    DNA fragmentation
    Annexin V
    Cytochrome c release
    Caspase-3
    U0126
    DN Raf
    DN ERK
    [77]
    Primary mouse kidney proximal tubular epithelial cells EGF deprivation 120 h Cell viability U0126
    PD98059
    [78]
    Primary human bone marrow stromal cells Leptin 12 h Cell viability
    Cytochrome c release
    Caspase-3
    U0126
    PD98059
    [79]
    Human leukemia U937 Bufalin 12 h DNA fragmentation MEK antisense [11]
    Pig renal tubular epithelial cells LLC·PK1 Escherichia coli toxin 44 h AnnexinV
    PARP cleavage
    PD98059 [80]
    Mouse monocyte/macrophage J774.2 Francisella tularensis infection 42 h Annexin V
    DNA fragmentation
    U0126
    PD98059
    [81]
    Human osteosarcoma cell line HOS and U2OS Chelerythrine 4 h DNA fragmentation
    Caspase-8, -9, -7
    PARP cleavage
    MEKCA U0126
    PD98059
    DN Ras
    [82]
    Mouse embryonary fibroblast RacN17 Cdc42N17 ND Membrane integrity PD98059 [83]
    Mouse embryonary fibroblast RacN17 Cdc42N17 24 h Annexin V PD98059
    DN DUSP6
    [84]
    Rat fibroblast Rat1 RAF-CAAX Constitutive DNA condensation RAF-CAAX [85]
    Mouse immortalized fibroblast NIH 3T3 DAPK Constitutive Cell viability MEKCA [87]
    Murine erythroleukemia
    DP16.1/p53ts cells
    P53 induction Constitutive Annexin V U0126
    Raf CA
    [86]
    Human breast adenocarcinoma MCF-7 ΔRAF1 Constitutive DNA fragmentation
    Vacuolization
    ΔRAF1 [88]
    Human primary osteoblast ΔRaf1
    HrasV12 T35S
    Constitutive Cell viability
    Membrane integrity
    Annexin V
    DNA fragmentation
    ΔRaf1
    HrasV12T35S
    [89]
    HEK293T IGF-I receptor 48 h Membrane integrity Caspase-3
    Vacuolization
    U0126
    MEK siRNA
    [90]
    HEK293 ΔRaf1:ER
    Anti-Fas(CH11)
    Constitutive Membrane integrity DNA fragmentation
    DNA condensation
    Annexin V
    Caspase-8, -3
    PARP cleavage
    Vacuolization
    ΔRaf1:ER
    U0126
    [91]
    Murine fibrosarcoma cells L929 TNFα Cell viability
    LC3-II induction
    Beclin induction
    U0126
    PD98059
    [48]
    Human breast adenocarcinoma MCF-7 TNFα 10 h Cell viability
    LC3-II induction
    U0126
    PD98059
    [49]
    Mouse RAW264.7 macrophages NO 2 h
    constitutive
    Cell viability
    Membrane integrity
    BNIP-3 induction
    U0126 [66]
    Transformed mouse
    mesengial MES-13 cells
    Cadmium 3 h Cell viability MTT
    LC3-II induction
    PD98059
    [69]
    Human colon adenocarcinoma HT29 Amino acid depletion 4 h GAIP phosphorylation
    Autophagic LDH sequestration
    PD98059 [114]
    Human colon adenocarcinoma HT29 Amino acid depletion
    ATA
    RasV12S35
    4 h
    constitutive
    Proteolysis
    LDH sequestration
    RasV12S35 [115]
    Human colon adenocarcinoma
    HCT-15
    Soyaponin 3 h MDC incorporation U0126 [116]
    Mouse 42GPA9 Sertoli cell line Lindane 24 h LC3 relocalization
    Vacuolization
    PD98059
    U0126
    [117]
    Primary human lung fibroblasts WI38 Dihydrocapsaicin 4 h LC3-II induction
    Caspase-3, -7
    LC3 relocalization
    PD98059
    ERK siRNA
    [118]
    Human OVCA-420 PEA-15 48 h Membrane integrity
    Acidic vacuoles
    ERK siRNA [119]
    Primary human fibroblast IMR90 RasV12 Constitutive LC3-II induction [120]

    The proapoptotic function of the Ras/Raf/ERK pathway is well documented for apoptosis induced by DNA-damaging agents, such as etoposide [12–15], doxorubicin [13,16–20], UV [13] and gamma irradiation [21]. ERK activity has been particularly implicated in cisplatin-mediated apoptosis in renal cells [15,22–32] (Table 1).

    ERK activity has also been involved in cell death induced by various antitumor compounds, such as resveratrol [33], quercetin [34], phenethyl isothiocyanate [35], betulinic acid [36], apigenin [37], oridonin [38], miltefosine [39], shikonin [40] or taxol [10,41] (Table 1).

    Most of these drugs induce the intrinsic apoptotic pathway. However, ERK activity has also been involved in activation of the extrinsic pathway by death receptor ligands such as TNF-related apoptosis-inducing ligand (TRAIL) [34,42–46], TNFα [47–49], Fas [50,51] or CD40 ligand [52,53]. Cell death induced by other death pathways that occur in response to zinc [54,55], oxidation [56–59], especially in response to ONOO [60], H2O2 [61–64] or NO treatment [65–67], toxic compounds such as cadmium, [17,68,69], benzo[a]pyrene [70], asbestos [71] or arsenic [17,72], also require ERK activity. Many death stimuli, such as estradiol [73] or its antagonist tamoxifen [74], interferon-α [75], cephalosporin [76], the calcium mobilizer calcimycin [77], epidermal growth factor (EGF) deprivation [78], leptin [79], bufalin [11], bacterial infection [80,81], chelerythrine [82] and the dominant negative form of Rac and Cdc42 [83,84], are sensitive to inhibition of the Ras/Raf/ERK pathway (Table 1).

    Conversely, constitutive activation of ERK by dominant active Raf1 combined with c-Myc expression [85] or p53 induction [86], or constitutively active MEK combined with death-associated protein kinase (DAPK) expression [87], could induce apoptosis without any other stimulus. Moreover, in rare cases, activation of the Raf/ERK pathway alone mediated by Raf1 [88,89], RasV12S35 [89], insulin-like growth factor type I (IGF-I) receptor [90] expression or by ΔRaf1:ER induction [91] was sufficient to promote cell death.

    Mechanisms of Ras/Raf/ERK-mediated apoptosis

    ERK activity has been associated with classical markers of apoptosis execution, such as effector caspase-3 activation, poly(ADP-ribose) polymerase (PARP) cleavage, annexin-V staining, DNA condensation or DNA fragmentation (see Table 1 for details). Depending on the cell type and the nature of the injury, activation of the Ras/Raf/ERK pathway is associated with the intrinsic apoptotic pathway characterized by the release of cytochrome c from mitochondria [22,38,57,77,79] and activation of initiator caspase-9 [20,27,29,57,59,82] or with the extrinsic apoptotic pathway, which relies on the activation of an initiator caspase-8 [27,28,40,42,68,82].

    ERK promotes caspase-8 signaling and activation

    The Ras/Raf/ERK pathway potentiates activation of death receptors by increasing the level of death ligands such as TNFα [28] or FasL [51] or death receptors such as Fas [39,89,91], DR4 [44,46] or DR5 [42,44,46]. ERK activity also promotes the induction of FADD, an adaptator of caspase-8 to the death receptors [39,91]. Because ERK-mediated caspase-8 activation requires de novo protein synthesis [68,91], it may reflect the activation of transcription factors regulated by ERK, such as c-Fos, which has been associated with the upregulation of DR4 and DR5 [44]. However, we have shown that ERK-induced caspase-8 activation could be independent of Fas and FADD upregulation, suggesting death receptor-independent modes of caspase-8 activation by ERK [91]. FADD bears a death effector domain that mediates caspase-8 activation. A very similar structure is found to bind ERK in vanishin [92] and PEA-15 [93], proteins that regulate both ERK and FADD activity [94]. These observations suggest that differential interactions between death effector domain-containing proteins that bind either ERK or caspase-8 could mediate death receptor-independent activation of the extrinsic pathway of apoptosis.

    The control of cytochrome c release by Bcl-2 family proteins

    As shown above, ERK activity is associated with DNA-damaging agents and antitumor compound-induced apoptosis, which are often described as inducing the intrinsic pathway of apoptosis. Therefore, several studies have suggested that the Ras/Raf/ERK pathway is involved in this pathway. Indeed, ERK activity has been shown to directly affect mitochondrial function by decreasing mitochondrial respiration [25,58] and mitochondrial membrane potential [58,63,79], which could lead to mitochondria membrane disruption and cytochrome c release [22,38,57,77,79]. Interestingly, active ERK has been found to be localized to mitochondrial membranes [25,58,63].

    ERK activity could also promote cytochrome c release by modulating Bcl-2 family protein expression. MEK/ERK activity has been associated with the upregulation of proapoptotic members of the Bcl-2 family, such as Bax [20,30,40,62,67,77], p53 upregulated modulator of apoptosis (PUMA) [20,86] and Bak [75], as well as the downregulation of antiapoptotic members, such as Bcl-2 [13,18,20,23,41,45,89] and Bcl-Xl [23,45]. In addition, ERK-activated caspase-8 induces the release of cytochrome c through proteolytic activation of the proapoptotic member Bid [45].

    ERK promotes p53 stability and activity

    The regulation of Bcl-2 family proteins has been tightly associated with transcriptional activity of the tumor suppressor gene p53. Apoptosis induced by DNA-damaging agents correlates with p53 upregulation and modulation of Bcl-2 family proteins in an MEK-dependent manner [13,18,20,23,24,33,40,41,48,77,95]. ERK-mediated p53 upregulation is associated with p53 phosphorylation on serine 15 [20,24,33,70,86,95,96], which stabilizes p53 protein and promotes its accumulation by inhibiting an association with Mdm2 [96]. This is supported by the ability of ERK to bind p53 [18,95] and to phosphorylate p53 on serine 15 in vitro [95,96]. Moreover, Mdm2 phosphorylation on serine 166, which is associated with its ubiquitin ligase activity toward p53, is inhibited upon sustained ERK activation [97]. ERK activity is implicated in p53 phosphorylation on threonine 55, promoting DNA-binding activity and Bcl-2 downregulation [18].

    c-Myc, which is stabilized by ERK through phosphorylation on serine 62, increases the proapoptotic functions of p53 [98]. Interestingly, when combined with c-Myc overexpression, constitutive activation of ERK is sufficient to induce apoptosis in Rat-1 cells [85] and to potentiate TRAIL-induced apoptosis in primary fibroblasts [43].

    The use of p53-deficient cells [35,83], p53 siRNA [70], p53 antisense [33,77], p53 inhibitor pifithrin-α [20,33,41], temperature-sensitive allele of p53 [86] or inducible p53 [24] showed that ERK-mediated p53 expression is required for apoptosis. However, in other studies, the Ras/Raf/ERK pathway is able to induce apoptosis independently of p53 [24,35,36,41].

    Together, these data suggest that upregulation of the tumor suppressor p53 may be an important mechanism of Ras/Raf/ERK-induced apoptosis.

    Other mediators of Ras/Raf/ERK-induced apoptosis

    Cytosolic phospholipase A2 (cPLA2) is a potential mediator of Ras/Raf/ERK pathway-induced apoptosis through intrinsic as well as extrinsic pathways. The Fas receptor in Sertoli cells [51], B cell receptor (BCR) in B lymphoma [99] or leptin in adipocytes [79], all promote MEK-dependent cPLA2 induction and activation. ERK can directly activate cPLA2 by phosphorylation at serine 505 [100,101]. The use of cPLA2 inhibitor AACOCF3 suggests that cPLA2 was necessary for ERK-induced apoptosis by a mechanism that promotes FasL induction [51] or cytochrome c release [79].

    Like death receptors and FADD, DAPK contains a death domain. ERK was shown to bind to DAPK and increase its catalytic activity by phosphorylation on serine 735 [87]. DAPK activation results in apoptosis due to cell detachment [87] or increase in TNF receptor function [47].

    In MCF-7 cells and primary osteoblasts, activation of Raf/ERK pathway-induced apoptosis was the consequence of cellular detachment from the matrix, which was in this case due to a decrease in integrin β1 expression [88,89].

    DNA-damaging agents have been shown to mediate sustained ERK activation through the protein kinase ataxia telangiectasia mutated [13,102,103].

    Implication of Ras/Raf/ERK pathway during apoptosis in vivo

    Following tissue injury

    ERK activity has been clearly implicated in neurodegenerative diseases and brain injury following ischemia/reperfusion in rodents (for a review see [6,104,105]). The Ras/Raf/ERK pathway also plays a key role in mouse models of acute renal failure induced by cisplatin [28] or lung injury induced by hyperoxia [57], as treatment with MEK inhibitors prevents apoptosis and tissue destruction in these models.

    During development

    Proapoptotic ERK activity has also been reported in developmental models. During germinal cell development, PEA-15, a cytoplasmic death domain-containing protein that binds and sequesters ERK, is highly expressed in the cytoplasm of Sertoli cells, spermatogonia and spermatocytes, inducing a cytoplasmic ERK localization. Interestingly, testis isolated from PEA-15-deficient mice display an abnormal nuclear accumulation of ERK in germinal cells, which correlates with increased apoptosis [106]. In Caenorhabditis elegans, loss-of-function alleles of lin-45 (RAF homolog), mek-2 (MEK homolog) and mpk-1 (ERK homolog), have presented genetic evidence for a direct role of the Ras/Raf/ERK pathway in germinal cell apoptosis [72]. In unfertilized eggs of starfishes Asterina pectinifera and Marthasterias glaciali, v-mos Moloney murine sarcoma viral oncogene homolog (MOS)-dependent sustained ERK activity led to protein synthesis-dependent synchronous apoptosis [107–109]. Moreover, maintaining ERK activity in fertilized eggs by MOS injection is sufficient to induce apoptosis [107]. During metamorphosis in ascidian Ciona intestinalis, sustained nuclear activity of ERK homolog (Ci-ERK) in the tail is required for the induction of apoptosis (caspase-3-like activity) and necessary for tail regression [110]. Finally, during limb development in chick embryos, ERK activity is inhibited in the mesenchyme by FGF8-induced DUSP6 activity. When activated by the expression of constitutively active MEK1, downregulation of DUSP6 or by the expression of a phosphatase-inactive mutant of DUSP6 (C294S), sustained ERK activity induces massive apoptosis and prevents limb development [111]. These results strongly indicate that Ras/Raf/ERK pathway-mediated apoptosis is not only associated with in vitro manipulation of cell lines, but also plays a key role in vivo during development and following tissue injury.

    The role of the Raf/ERK pathway in the induction of cell death should not be restricted to apoptosis, i.e. caspase-dependent cell death. In some cases, the methods used to assess cell viability, based on cell metabolism or membrane permeabilization measurements (see Table 1), cannot distinguish between apoptosis and other forms of cell death, such as necrosis or autophagy.

    Cytoplasmic vacuolization: lysosomal cell death and autophagy

    Autophagy is a genetically regulated program, initially identified as a cell survival mechanism to protect from nutrient deprivation. However, in certain conditions, autophagy results in a form of cell death now described as type II programmed cell death [112].

    We and others have shown that constitutive activation of ERK by active Raf [88,91], cadmium [68] or IGF-I receptor [90] induced a form of cell death that correlated with extensive cell rounding and the formation of cytoplasmic macrovacuoles, which pushed the nucleus and the cytoplasm to the side of the dying cell. Although cell death was associated with caspase-8 activation [68,91], this massive vacuolization is unrelated to the classical features of apoptosis. This morphology could be a sign of autophagic programmed cell death, but also of paraptosis, a form of caspase-independent cell death associated with cytoplasmic vacuolization [90]. Interestingly, other studies using cadmium [69,113] or TNFα treatment [49] have clearly associated ERK activation with autophagic programmed cell death rather than with apoptosis. This is supported by several studies that have associated ERK activity with neuron autophagic cell death in the course of a neurodegenerative disease [6,104,105]. In addition, ERK activity has also been associated with autophagy and autophagic cell death in many non-neuronal cellular models (see Table 1) in response to different stresses, such as amino acid depletion [114] and aurintricarboxylic acid (ATA) [115] in human colorectal cancer cell line HT29, soyasaponins [116] in human colon adenocarcinoma HCT-15, lindane [117] in the mouse Sertoli cell line, dihydrocapsaicin [118] in WI38 lung fibroblasts, cadmium in mesengial MES-13 cells [69,113] and TNFα treatment in MCF-7 [49] and L929 cells [48]. Interestingly, in human ovarian cancer cells, cytoplasmic sequestration of ERK by PEA-15 has been shown to promote autophagy [119]. Moreover, direct ERK activation by overexpression of constitutively active MEK can promote autophagy without any other stimulus [117].

    ERK-dependent autophagic activity is associated with classical markers of autophagy, such as induction of LC3 and conversion of LC3-I to LC3-II [48,49,118], induction of beclin 1 [48], induction of BNIP-3 [66] and activation of G-interacting protein (GAIP) by phosphorylation on serine 151 [114]. p53 is also associated with the autophagic process, as ERK-mediated phosphorylation of p53 on serine 392 [48] was involved in TNFα-induced autophagy.

    The lysosomal compartment plays an important role in autophagy by fusing with autolysosome vacuoles. In NIH3T3 and in human colon carcinoma HCT-116 cells, oncogenic forms of Ras, respectively, v-H-Ras and K-Ras, lead to increased sensitivity to the lysosomal cell death pathway induced by cisplatin or etoposide in an MEK-dependent manner. In these models, constitutive ERK activation leads to a decrease in the levels of lysosome-associated membrane protein-1 and -2 due to the induction and activation of cysteine–cathepsin B [15].

    In humans, ERK activity is potentially associated with limb sporadic inclusion body myositis, a disease characterized by cytoplasm vacuolization of muscle fibers. Interestingly, immunostaining of muscle samples from patients revealed a strong ERK accumulation in cytoplasmic vacuoles [120].

    Together, these results suggest that the Ras/Raf/ERK pathway can mediate autophagic type II programmed cell death.

    ERK-induced cytoplasm vacuolization associated with autophagy has some similarity with senescence-associated vacuoles. Interestingly, autophagy has recently been reported to be required for the efficient establishment of senescence induced by a constitutively active form of Ras or MEK [121].

    The Ras/Raf/ERK pathway and senescence

    Cellular senescence is an irreversible form of cell cycle arrest that prevents proliferation of damaged cells or cells that have surpassed their capacity to proliferate. In response to oncogenic hyperproliferative signals, primary cells undergo cell cycle arrest leading to premature oncogene-induced senescence [122]. Although aberrant activation of the Ras/Raf/ERK pathway promotes oncogenic transformation of immortalized cells, it is also tightly associated with senescence of primary cells. In human and rodent primary fibroblastic and melanocytic cells, senescence is triggered by constitutively active forms of Ras [123–126], PAK4 [125,127], Raf [97,126,128–133] or MEK [123,124,132,134]. This process is often prevented by use of MEK inhibitors (See Table 2).

    Table 2. Models of ERK-mediated senescence.
    In vivo/cellular model Induction of ERK activity Markers of senescence MEK activity Reference
    Primary human fibroblastic IMR90 Ras V12
    MEKCAQ56P
    β-galactosidase activity
    p21 p53 and p16/INK4A
    PD98059 [123]
    Primary human fibroblastic BJ cells RasV12
    MEKCA
    β-galactosidase activity
    p16/INK4A
    U0126 [124]
    Primary mouse fibroblastic cells Ras V12
    PAK4
    β-galactosidase activity
    p21, p19/ARF and p16/INK4A
    PD98059 [125]
    Mouse immortalized NIH 3T3 fibroblast ΔRaf1:ER p21 [128]
    Primary human fibroblast IMR90 ΔRaf1:ER β-galactosidase activity
    p21 and p16/INK4A
    PD98059 [129]
    Human prostate cancer LNCaP cells ΔRaf1:ER β-galactosidase activity
    p21
    [130]
    Mouse embryonic fibroblast ΔRaf1:ER p21, p53 and p19/Arf [131]
    Primary human fibroblast IMR90 Raf-CAAX
    ΔRaf1:ER
    β-galactosidase activity [97]
    Primary human melanocytes
    Primary human fibroblast BJ
    In vivo/human naevi
    B-Raf V600E β-galactosidase activity
    p21, p53 and p16/INK4A
    Heterochromatic foci
    [127]
    Primary human melanocytes
    In vivo/human Naevi
    B-Raf V600E β-galactosidase activity
    Heterochromatic foci
    p16/INK4A
    [135]
    In vivo/mouse lung tumor model B-Raf V600E p19/ARF Dec1
    Heterochromatic foci
    [133]
    Primary human melanocytes B-Raf V600E
    MEKCAQ56P
    β-galactosidase activity
    p16/INK4A, γH2AX
    Heterochromatic foci
    [132]
    Primary human melanocytes B-Raf V600E
    NRasQ61R
    β-galactosidase activity
    p53 and p16/INK4A
    [126]
    Primary human intestinal epithelial cells MEKCA SS218/222DD β-galactosidase activity
    p21 p53 and p16/INK4A
    [134]
    Primary human fibroblast WI38 DUSP4 (MKP-2) β-galactosidase activity [136]

    Mechanisms of ERK-induced senescence

    Ras/Raf/ERK pathway-induced senescence correlates with increased β-galactosidase activity and induction of classical senescence-associated genes, such as p16/INK4A, p53, p21 and p14-p19/ARF (Table 2), senescence-associated heterochromatic foci [127,132] and DNA damage foci [132]. In human primary fibroblasts IMR90, when senescence is provoked by inducible activation of Raf1:ER, it correlates with inhibition of AKT and dephosphorylation of Mdm2, which lead to p53 accumulation and growth arrest [97]. In the case of human BJ foreskin primary fibroblasts, senescence induced by ectopic expression of RasV12 or constitutively activated forms of MEK (MEKCA) requires ERK-induced p38 activation [124]. Interestingly, a phenotypic comparison between RasV12-, RasR61- and B-RafE600-induced senescence in human melanocytes suggests that the senescence programs are different: Ras-induced senescence was faster and was associated with massive cytoplasmic vacuolization, whereas B-Raf-expressing cells exhibited a more rounded morphology [126,132]. However, human and rodent cells induce different senescence programs. The use of mouse embryonic fibroblasts derived from knock-out mouse models suggested that Ras/Raf/ERK pathway-induced senescence relies on the induction of cell cycle regulators, such as p16/INK4A [123,125], p21 [126], p53 [123], p19/ARF [125]. In human primary fibroblasts, however, knock-down or inhibition of either p16 [127] or p53 [126] was not sufficient to reverse senescence, suggesting that these genes may have a redundant function controlling human senescence. Oncogene-induced senescence prevents transformation of human primary cells unless overridden by the presence of a cooperating oncogene, such as Myc. Indeed, overexpression of c-Myc in normal human melanocytes suppressed B-Raf- or N-Ras-induced senescence [126]. Myc expression is then continuously required for transformation, as downregulation of c-Myc in tumor-derived melanoma cells was shown to induce senescence [126].

    Senescence and subcellular ERK localization

    ERK-induced senescence has been associated with an aberrant control of its spatial activity. Kim et al. [135] found that reactive oxygen species (ROS) produced during senescence of human primary fibroblasts inactivate the cytosolic ERK phosphatase DUSP6, resulting in cytoplasmic sequestration of active ERK. However, other studies have suggested that senescence could also be the result of inhibition of nuclear ERK activity due to an increase in nuclear DUSP4 activity [136,137]. Thus, coordinate gene expression induced by nuclear ERK might be required to prevent the completion of a senescence program induced by increased cytoplasmic ERK activity.

    Implication of the Ras/Raf/ERK pathway in senescence in vivo

    Senescence induced by ectopic expression of an oncogene might reflect an artificially high expression level, as discussed in the study by Tuveson et al. [138] of the oncogene KRasD12. However, several results based on the expression of B-RafE600 at physiological level under the control of its own promoter, support the idea that Ras/Raf/ERK pathway-induced senescence is a physiological cellular response. For instance, in humans, naevi (moles) can be considered as an in vivo example of B-Raf-driven senescence. Naevi are melanocyte-derived benign tumors restrained from malignant progression by engagement of senescence. Naevi frequently harbor the oncogenic B-RafE600 or NRasR61 mutation [139] (which promote senescence of primary melanocytes [126,127,132]) and display markers of senescence such as β-galactosidase activity and high p16 expression [127,140]. Recently, a mouse model of inducible tumorigenesis in lung epithelium driven by the B-RafE600 oncogene revealed that expression of B-RafE600 alone was not sufficient to promote a severe tumoral phenotype, leading instead to benign hyperplastic lesions undergoing senescence-associated growth arrest [133]. In this model, p53 invalidation was necessary to promote B-RafE600-mediated transformation and malignant tumor formation [133].

    The hallmarks of ERK-mediated cell death: sustained and sequestered activity

    ROS as mediators of ERK-induced cell death

    In the majority of the studies related to cell death induced by the Ras/Raf/ERK pathway, ERK activation is unusually prolonged, i.e. ERK is maintained phosphorylated for between 6 and 72 h (see Table 1). Moreover, delayed treatments with U0126, a MEK inhibitor, have revealed that ERK activity is continuously required to induce cell death [91]. Despite constitutive activation of the pathway by oncogenes, levels of ERK phosphorylation in tumor cells are very variable [141], presumably due to phosphatase-driven feedback mechanisms. Because ERK-specific phosphatases are sensitive to ROS, we speculate that the main cause of sustained ERK activation is the presence of ROS, perhaps reflecting the levels of ROS scavengers in each particular model. The use of different ROS inhibitors demonstrated that ERK activation requires ROS production to induce cell death [14,54,56,57,60,61,65,76,78]. Indeed, chemical oxidants, such as H2O2, peroxynitrite ONOO or NO (see Table 1), induce ERK, whereas many stimuli implicating ERK in cell death promote the production of ROS [14,35,54,55,71,76,82]. Moreover, DNA-damaging agents, such as doxorubicin, cisplatin or etoposide, catalyze the formation of ROS [142]. In addition, ERK activity could be directly responsible for ROS production by upregulating inducible NO synthase [80]. Thus, ROS-mediated prolonged ERK activation might be the crucial mechanism implicating the functions of the Ras/Raf/ERK pathway in cell death.

    ROS promote sustained ERK activation

    Mechanisms of ROS-mediated ERK activation upstream of ERK

    ROS can stimulate the Ras/Raf/ERK pathway by promoting the activation of tyrosine kinase receptors, such as platelet-defined growth factor receptor or EGF receptor [26,60,61], and adaptor proteins, such as Shc [143]. ROS can also increase signaling by direct oxidation of residue C118 on Ras, a reaction that potentiates recruitment and activation of Raf at the plasma membrane [144,145]. Other proteins implicated in Raf activation, such as Src [61], protein kinase C (PKC)-δ [25,29,61] or the cGMP pathway [146], could also be activated by ROS. Moreover, direct oxidation of cysteine residues in the cystein-rich domain of Raf promotes its autoactivation [147]. Downstream of Raf, peroxynitrite ONOO can also cause nitration and autophosphorylation of MEK [61].

    ROS and inhibition of ERK phosphatases

    Finally, ERK activity could be prolonged through the inhibition of tyrosine phosphatases and DUSP by ROS. Indeed, enzymatic activity of DUSP and tyrosine phosphatases requires a catalytic cysteine residue sensitive to oxidation [3,4]. ROS have been shown to inhibit ERK-directed phosphatases, DUSP1 and DUSP6, by oxidation of their catalytic cysteine residues, C258 and C293, respectively [148,149], as well as ERK tyrosine phosphatases PP1/2A by oxidation of their conserved catalytic residue C62 [135]. Thus, the control of phosphatases that downregulate ERK activity plays a crucial role in the outcome of Ras/Raf/ERK pathway signaling.

    Together, these data indicate that ROS can initiate and sustain ERK activation by different mechanisms. Interestingly, cell death has been associated with a biphasic activation of ERK, which could reflect this dual control of ERK activity by ROS [62,68].

    The importance of subcellular localization of ERK activity

    In most cases, prolonged ERK activation alone, such as in models expressing constitutively active forms of upstream kinases, is not sufficient to promote cell death [15,63,64,66,82,85–87]. In normal cells, subcellular localization of ERK is tightly regulated by scaffold proteins and docking phosphatases that allow nuclear accumulation of dephosphorylated ERK to terminate signaling [2]. Thus, in addition to a sustained ERK activity, the outcome of ERK-mediated cell death might also rely on an aberrant subcellular localization. Indeed, apoptosis induced by estradiol [73], tamoxifen [74], zinc [54,55] cephaloridine [76], doxorubicin [20], revestratol [33] or dominant negative mutant of Rac or Cdc42 [84] correlated with sustained nuclear ERK activity. As mentioned previously, nuclear activation of ERK is also associated with apoptosis during Ciona intestinalis development [106] and in mouse testis deficient for PEA-15 [106]. Interestingly, some of the compounds that induce nuclear ERK activity are associated with the production of ROS [54,55,76], which could promote nuclear accumulation of active ERK due to inhibition of DUSP. In MDA-MB-231 human breast cancer cells, taxol-induced apoptosis was abrogated by induction of nuclear DUSP1. In this study, DUSP1 induction clearly inhibited both ERK and JNK activity [150]. Because nuclear DUSPs (especially DUSP1 and -4) also control JNK and p38 phosphorylation [3,4], any modification of DUSP activity or expression could also increase cell death by activation of the stress pathways.

    Cytoplasmic sequestration of ERK has also been associated with different forms of cell death. Cytoplasmic sequestration of ERK by binding to PEA-15 promotes autophagy [119], whereas sustained cytoplasmic ERK activity induces senescence in human primary fibroblasts [135–137]. Together, these data suggest that sustained activation of ERK in different subcellular compartments is not tolerated and results in different forms of cell death (see Fig. 1).

    Details are in the caption following the image

    The hallmarks of ERK-mediated cell death: sustained and sequestered activity. ERK activity induces the expression of many genes, including its own regulators, the DUSP (ERK-specific phosphatases). Thus, ERK activity rapidly reaches a steady state and its death-promoting activity remains at low levels (gray line). Any agent that provokes a sustained activation of ERK, such as ROS that inhibit DUSP, induces the progressive accumulation of death-promoting factors up to a level that induces cell death. The activation of ERK (arrow) might also transiently increase death-promoting activities of other death stimuli, such as chemotherapeutic agents. In addition to DUSP inhibition by ROS, ERK-mediated cell death is characterized by a deregulation of subcellular active ERK localization. Sustained cytoplasmic ERK activity might promote senescence or autophagy, whereas sustained nuclear sequestration of ERK activity might trigger apoptosis. In both conditions, sequestration of ERK depends on subcellular anchors, such as PEA-15, in the cytoplasm.

    The limits of ERK1/2-mediated cell death studies, the specificity of MEK inhibitors

    In many studies, implication of the Ras/Raf/ERK pathway in the induction of cell death is based uniquely on the sole use of MEK1/2 inhibitors PD05059 or U0126 (see Table 1). The weakness of these inhibitors is that they inhibit both MEK1/2 and MEK5 [151]. Interestingly, it has recently been shown that constitutive activation of the MEK5/ERK5 pathway could promote apoptosis of meduloblastoma cells [152] or thymocytes [153,154], through Nur77-dependent mechanisms [153,154]. As a consequence, in those types of cell, some of the effect attributed to ERK1/2 might also be caused by the MEK5/ERK5 pathway. The use of PD184352, a more recent MEK1/2 inhibitor that does not target MEK5 [155], could help to distinguish between the effects of MEK1/2 and MEK5 in those cells.

    Conclusions

    Together, these data clearly demonstrate that the Ras/Raf/ERK pathway plays a critical role in promoting several forms of cell death in response to numerous stress stimuli both in vitro, with various cellular models, and in vivo. A common hallmark of this response is the sustained activation of ERK, which contrasts with the transient nature of ERK stimulation found in situations where ERK regulates other cell fates. As depicted in Fig. 1, ERK activates its own phosphatases, inducing a feedback loop that, within hours, restores a basal level of ERK activity. At least in the cellular models depicted in Table 1, ERK stimulation induces the expression of gene products with death-promoting activity. We can speculate that the feedback loop decreasing subcellular ERK activity over time prevents these death-promoting factors reaching a threshold concentration that triggers cell death. Consequently, any agent affecting the kinetics of ERK activity in a given cellular compartment (such as the ROS that inhibit DUSP in Fig. 1) has a potential to induce cell death. Given the importance of the spatiotemporal regulation of ERK activity for the control of cell division [2], the induction of cell death could be seen as negative feedback mechanism preventing uncontrolled cell proliferation.

    The Ras/Raf/ERK pathway is among the most commonly deregulated pathways identified in tumors, as indicated by frequently observed activating mutations in Ras or B-Raf oncogenes. Thus, this pathway is currently the target of new antitumor strategies, based on the inhibition of upstream ERK regulators. However, because ERK activation is implicated in DNA-damaging agent-induced cell death (see Table 1), inhibiting ERK activity in combination therapy with classical antitumor compounds, such as cisplatin or doxorubicin, might affect the efficiency of such compounds.

    Because prolonged ERK activation has been shown to promote the death of human cancer cell lines from different origins (see Table 1), this property of the Ras/Raf/ERK pathway to induce cell death could be used to target cancer cells. However, although tumor cells escape Ras/Raf/ERK pathway-induced senescence by inactivating effectors of senescence, such as p53 or p16/INK4A, mechanisms involved in ERK-induced cell death might also be silenced in tumor cells. Tumor cells with high ERK activity might also have re-modeled the ERK signaling to escape ERK-mediated cell death. Thus, the crucial biochemical events underlying sensitivity or resistance to ERK-mediated cell death remain to be fully understood. We propose the hypothesis that in tumor cells harboring strong ERK activity, the alteration of compensating pathways (PI3K/AKT, Wnt, etc…) would unleash the cell killing ability of ERK. Alternatively, if reagents were able to sequester ERK in a given subcellular compartment, the changes in the spatiotemporal regulation of ERK might be lethal. Moreover, because sustained ERK activity is required to promote cell death, such strategies would only target cancer cells with deregulated ERK activity and not normal cells in which ERK activation is transient.

    Acknowledgements

    S. Cagnol is supported by the Canadian Institutes for Health Research grant CIHR MT-14405. We thank Dr Brendan Bell for careful reading of the manuscript. [Correction added on 30 October 2009 after first online publication: in the preceding sentence the name ‘Dr Brendan Bell’ was corrected.]