Open access peer-reviewed chapter

Regulation of MAPK ERK1/2 Signaling by Phosphorylation: Implications in Physiological and Pathological Contexts

Written By

Dadnover Vargas-Ibarra, Mariana Velez-Vasquez and Maria Bermudez-Munoz

Submitted: 14 October 2020 Reviewed: 06 March 2021 Published: 12 April 2021

DOI: 10.5772/intechopen.97061

From the Edited Volume

Post-Translational Modifications in Cellular Functions and Diseases

Edited by Shibo Ying

Chapter metrics overview

441 Chapter Downloads

View Full Metrics

Abstract

Protein phosphorylation represents a rapid and reversible post-translational regulation that enables a fast control of protein activation that play key roles in cell signaling. For instance, Mitogen Activated Protein Kinase (MAPK) pathways are activated upon sequential phosphorylations, resulting in phosphorylation of cytosol and nuclear targets. We focus here on MAPK ERK1/2 signaling that accounts for diverse cellular responses such as cell cycle progression, proliferation, differentiation, senescence, migration, formation of GAP junctions, cell adhesion, cell motility, survival and apoptosis. We review the role of protein phosphorylation in MAPK ERK1/2 activation, in its regulation in time and space and how its dysregulation can lead to tumorigenesis.

Keywords

  • phosphorylation
  • cell signaling
  • MAPK
  • ERK1/2
  • kinase
  • phosphatase
  • cancer
  • inhibitors

1. Introduction: cell signaling regulation by phosphorylation

Among post-translational modifications, protein phosphorylation is the most common. Vitellin was the first protein which phosphorylation was discovered, by Phoebus Levene in 1906 [1, 2]. In 1954, Burnett and Kennedy reported the process of enzymatic phosphorylation. Then, Edwin Krebs and Edmond Fischer described how phosphorylation and dephosphorylation can take place and they demonstrated how the process is governed by enzymes [3, 4]. In 1992, the Nobel Prize in Physiology or Medicine was awarded jointly to Edmond H. Fischer and Edwin G. Krebs for their discoveries concerning reversible protein phosphorylation as a biological regulatory mechanism.

Phosphorylation is a reversible protein modification and results from the addition of a phosphate group (PO4) to the polar group of amino acids. The most common amino acids that are phosphorylated are serine (Ser), threonine (Thr) and tyrosine (Tyr). Although phosphorylation of histidine and aspartate residues can also occur, they are less stable than others. Phosphorylation of a protein can change binding to other proteins: because each phosphate group has two negative charges, phosphorylation can cause a conformational change in the protein by attracting a cluster of positively charged amino acid side chains. This can change the binding of ligands on the protein surface and therefore its activity. On the other hand, the addition of a phosphate group to a protein can be recognized by other proteins having for instance SH2 and PTB domains, that then can attach to phosphorylated proteins such as the cytoplasmatic tail of receptor tyrosine kinases (RTK). Finally, phosphorylation can mask a binding site that otherwise holds two proteins together and then can disrupt this interaction.

Enzymes that catalyze the addition of a phosphate group to a protein are kinases; the reaction is unidirectional because of the large amount of free energy released when the phosphates bonds are broken in ATP to produce ADP. The human genome includes more than 500 protein kinases, and it is estimated that more than one-third of the 10,000 proteins in a typical mammalian cell are phosphorylated at any given time, many with more than one phosphate. Conversely, phosphatases are enzymes that remove a phosphate group from a protein, having the opposite function of kinases. Dephosphorylation has more rapid kinetics than phosphorylation by kinases. The human genome contains more than 200 phosphatases, classified into different families including protein tyrosine phosphatases (PTP), the metal-dependent protein phosphatase PPM, the phosphoprotein phosphatase (PPP) that are pSer/pThr- specific, the dual specificity phosphatase (DUSP) family and the PTEN family of lipid phosphatases [5].

Protein phosphorylation may occur at a single site that primes location for subsequent phosphorylations or directly at multiples sites. Thus, a single protein kinase or multiple kinases may act on the target protein, creating a synchronized cascade of phosphorylations. These events participate in dynamic intracellular signaling that enable cells to respond to extracellular stimuli and to adapt to internal changes. Mitogen-protein activated kinases (MAPK) are conserved kinases in eukaryotes, integrating cell signaling pathways that regulate processes such as cell proliferation, cell differentiation and cell death, from yeast to humans. There are four independent MAPK pathways: MAPK ERK1/2, ERK-5 (also referred to as BMK-1), c-Jun N- terminal kinase (JNK), and p38 signaling families. MAPK modules contain 3-tier kinases that are sequentially activated by phosphorylation. MAPK proteins are designated from upstream to downstream signaling pathway: MAPK kinase kinase (MAPKKK) phosphorylates MAPK kinase (MAPKK); MAPKK phosphorylates and thus activates MAPK. We will focus on MAPK ERK1/2 signaling to illustrate how a particular post-translational modification such as phosphorylation can regulate a signaling pathway and how its dysregulation can be implicated in pathological processes such as tumorigenesis.

Advertisement

2. MAPK ERK1/2 pathway: a cell signaling of sequential phosphorylations

The Extracellular Signal-Regulated Kinases (ERK) have key roles in processes like cell growth, cell proliferation and cell survival. In humans, there are three isoforms of ERK: ERK-1, ERK-2 and ERK-5. Hereon we will concentrate on classical MAPK ERK1/2 to comprehend how this signaling is regulated by phosphorylation.

In the canonical human MAPK ERK1/2 pathway there are three types of MAPKKK (A-Raf, B-Raf and Raf-1 or C-Raf kinases), two MAPKK (MEK1, MEK2) and two MAPK ERK-1, ERK-2. Interestingly, MAPK ERK1/2 signaling is basically regulated by phosphorylations. On the first level, Raf are serine/threonine-protein kinases that phosphorylate human MEK on Ser-218 and Ser-222, producing their activation. The Raf family of kinases includes three isoforms with high homology and a similar domain organization. On the second level, MEK1/2 are dual specificity protein kinases that phosphorylate a threonine and a tyrosine residue in a Thr-Glu-Tyr sequence located in ERK1/2, rendering them active. While human ERK-1 is phosphorylated on Thr-202 and Tyr-204, ERK-2 is phosphorylated on Thr-185 and Tyr-187 residues for activation. Phosphorylation of ERK1/2 by MEK1/2 leads to the rearrangement of several polar contacts, which results in conformational changes in neighboring structural elements (reviewed on [6]). Finally, ERK1/2 are serine/threonine kinases that phosphorylate a wide variety of substrates in different subcellular compartments including the Golgi apparatus, the mitochondrial membrane, the cytoplasm and the nucleus.

MAPK ERK1/2 phosphorylate substrates in a short Pro-X- Ser/Thr-Pro consensus motif (X representing any residue) and interactions with docking sites are important for specificity. Two motifs have been described, the D- and F-motifs, that can cooperate to enhance the substrate affinity of ERK and to set phosphorylation kinetics [7]. ERK1/2 phosphorylate more than 600 proteins, leading to responses such as cell cycle progression, proliferation, cytokinesis, transcription, differentiation, senescence, cell death, migration, formation of GAP junctions, actin and microtubule networks, neurite extension, cell adhesion and motility, survival and apoptosis [8]. To ensure that these cell responses are adaptive to stimuli in space and time, a fine regulation of MAPK signaling is thus necessary. Remarkably, control of ERK1/2 signaling is in part ensured by phosphorylations and dephosphorylations.

Advertisement

3. Regulation of MAPK ERK1/2 by kinases and phosphatases

The MAPK ERK1/2 has at least 3 tiers of regulation: the control of the phosphorylation and thus of the activity of Raf, MEK1/2 and ERK1/2. Additionally, upstream signals from cell receptors to Raf and downstream targets of ERK1/2 play an active role in regulating the MAPK ERK1/2 pathway. Interestingly, mechanisms of MAPK regulation are based partly on the same mechanisms that activate this pathway: phosphorylation events. In this section we specify how phosphorylation can regulate MAPK ERK1/2 signaling from Raf to ERK1/2 by the activity of kinases in feedback signals, and through dephosphorylation by phosphatases.

When RTK are activated by growth factors, their phosphorylated tyrosines enable the coupling of adaptor proteins such as GRB2. This latter binds with SOS, a guanine exchange factor that promotes the activation of Ras. From this level, MAPK ERK1/2 signaling axis exerts feedback regulations through phosphorylations. Growth factor stimulation (like epidermal growth factor EGF) of the cell induces the phosphorylation of four serine residues in a region encompassing three proline-rich SH3-binding sites in the C-terminal domain of SOS1 [9]. These phosphorylation events are realized by ERK1/2 and constitute a negative feedback regulation that leads to a reduction in Ras activation. Kinetic simulation model using parameters collected in living cells found that possibly more than four phosphorylation sites decisively suppress SOS activity [10]. Indeed, SOS1 is also phosphorylated by the ERK1/2 effector ribosomal S6 kinase 2 (RSK-2) on Ser1134 and Ser1161, leading to the recruitment of 14-3-3 and is thus a negative regulation of ERK1/2 activity [11] (Figure 1 and Table 1).

Figure 1.

Representative phosphorylation events leading to activation and feedback signaling in the MAPK ERK1/2 pathway. Phosphorylation constitutes activation (red arrows) or inhibition (black arrows) of proteins of MAPK ERK1/2 signaling. Specific details are provided in Table 1.

Number in figureProtein phosphorylatedPhosphorylation siteKinaseType of feedbackConsequenceReferences
1SOS1Ser1132, Ser1167, Ser 1178, Ser 1193ERK1/2NegativeDecreased binding affinity of Grb2 to human Sos1[9]
2SOS1Ser1134, Ser1161RSKNegativeFacilitates 14-3-3 binding, decreasing MAPK activation[10]
3RasGRP2Ser394ERK1/2NegativeInhibits RasGRP2 ability to activate Rap1, leading to decreased activation of ERK1/2[12]
4MEK1/2Proline-rich sequence (PRS)PAK1PositiveEnhances MEK1/2 interaction with C-Raf[13, 14, 15]
5KSR1Thr260, Thr274, Ser443ERK1/2NegativeInterrupts association of KSR1 with B-Raf and MEK1/2, driving the release of KSR1 from the plasma membrane[16, 17]
6MEK1Thr292ERK1/2NegativeInhibits MEK1 kinase activity towards ERK1/2, interferes with the binding of MEK1 to ERK2 and reduces the ability of PAK to phosphorylate MEK1 on S298 (required for the activation of MEK1 by cell adhesion)[18, 19]
7C-RafSer29, Ser43, Ser642, Ser289, Ser296, Ser301ERK1/2NegativeDesensitized C-Raf, do not localize to the plasma membrane and do not engage with activated Ras[20]
8DUSP6Ser159, Ser174, Ser197ERK1/2NegativeInduces degradation of DUSP6[21, 22]
9B-RafSer445Rap1PositiveActivation of B-Raf[23]

Table 1.

Feedback phosphorylation events in MAPK ERK1/2 pathway.

In platelets and nexus ERK1/2 is also activated downstream of the small GTPase Rap1. RasGRP2 is the predominant guanine exchange factor that specifically activates Rap1. RasGRP2, playing a similar role to SOS for Ras, is phosphorylated by ERK1/2 on Ser394 located in the linker region implicated in its autoinhibition. In this case, RasGRP2 phosphorylation results also in a negative feedback loop that determines the amplitude and duration of active ERK1/2 [12]. Moreover, Rap1 is able to phosphorylate and activate B-Raf (but not C-Raf) [23]. Upon cell adhesion and downstream of the small GTPase Rac, the serine/threonine-protein kinase PAK1 phosphorylates the MEK proline-rich sequence (PRS), enhancing its interaction with C-Raf [13] (Figure 1 and Table 1).

Regarding Raf, it has been shown that mitogenic stimuli induce the phosphorylation of C-Raf by ERK1/2 on six residues, needing MEK signaling. Hyperphosphorylation of these sites promotes the subsequent dephosphorylation of C-Raf by PP2A and the return to the inactive state [20]. On the other side, Raf interaction with MEK is also regulated by the inhibitor protein RKIP, which binds to both proteins preventing their physical association. RKIP interferes with the phosphorylation of MEK when bound to C-Raf. Association of RKIP with C-Raf is regulated partly by phosphorylation: phosphorylation of RKIP on serine 153 by PKC or putatively by ERK induces its dissociation from C-Raf [24, 25]. RKIP has then an important role in generating a switch-like behavior of MEK1/2 activity [26].

MEK1/2 is also the target of feedback regulation in the ERK1/2 pathway. Indeed, ERK1/2 phosphorylates MEK1 on Thr292, Thr286 and Thr386, resulting in reduced MEK activity and thus constitutes a negative feedback for MAPK ERK1/2 signaling [18, 19]. Moreover, MEK1 phosphorylation on Thr292 by ERK1/2 interferes with MEK1 binding to ERK2 and reduces MEK1 phosphorylation on S298 by PAK, required for the activation of MEK1 by cell adhesion [13, 14, 15] (Figure 1 and Table 1).

Another example of feedback regulation of MAPK ERK1/2 signaling by phosphorylation is the case of the protein scaffold KSR1. In fact, KSR1 can be phosphorylated in Thr256, Thr260, Thr274, Ser320, Ser443, Ser463 by ERK1/2 in vitro and depends on MEK1/2 activity. These KSR1 phosphorylations interrupt its association with B-Raf and MEK1/2, drive the release of KSR1 from the plasma membrane, representing then a negative feedback of MAPK ERK1/2 activation [16, 17] (Figure 1 and Table 1).

Another regulation of MAPK activity is accomplished by phosphatases that modulate later phases of ERK1/2 signaling. Ser/Thr phosphatases, protein tyrosine phosphatase and dual-specificity Thr/Tyr phosphatases (DUSP) dephosphorylate and thus inactivate ERK1/2. MAP Kinase Phosphatases (MKP) belong to DUSP and represent specific phosphatases that principally regulate MAPK activity in mammalian cells and tissues. While some DUSP dephosphorylate p38, JNK and ERK1/2, others are specific for p38/JNK or for ERK1/2. In this latter case are found cytoplasmic DUSP that inactivate ERK1/2 in the cytoplasm and include DUSP6/MKP-3, a specific phosphatase that binds to ERK1 and ERK2, inactivating them. This specificity is ensured by the fact that the interaction of DUSP6 with ERK1/2 is a requirement for the catalytic activation of the phosphatase through conformational changes [27, 28]. Interestingly, whilst inactivating ERK1/2, DUSP6 is in turn regulated by ERK1/2. Indeed, stimulation with serum or PDGF-B alone can induce a MEK-dependent phosphorylation of DUSP6 on Ser159, Ser174, and Ser197, which is followed by the degradation of the phosphatase by the proteasome [21, 22]. We have shown that another pathway involved in growth factor signaling, the PI3K/mTOR signaling pathway, accounts for a part of the phosphorylation and degradation of DUSP6 induced by serum growth factors. Furthermore, specific agonists of the mTOR pathway, such as amino acids or insulin/IGF-1 are also able to induce the phosphorylation and degradation of DUSP6. Mutagenesis studies identified Ser159 within DUSP6 as the target of the mTOR pathway [29]. Thus, DUSP6 is a point for double MAPK control: the phosphatase exerts a negative regulation for ERK1/2 activity but at the same time, ERK1/2 is able to phosphorylate DUSP6 and then induces its degradation. DUSP6 appears therefore as a spot for fine ERK1/2 signaling regulation in time. Moreover, DUSP6 is a branch-point for the crosstalk between two major signaling pathways induced by growth factors, the MEK/ERK1/2 pathway and the PI3K/mTOR pathway. Notably, both pathways are frequently overactivated in cancer cells. Thus, a regulation of MAPK ERK1/2 signaling in time and space is necessary to warrant cell physiological responses and to avoid aberrant signaling activation that facilitates pathological conditions.

Advertisement

4. Implications of phosphorylation in MAPK ERK1/2 regulation in time and space

MAPK ERK1/2 signaling can determine excluding cell responses such as proliferation and differentiation. Differences in cell responses upon MAPK ERK signaling depend on the regulation of the pathway through protein interactions by scaffolds and through inhibitory and adaptor proteins that enhance, decrease, or redirect the flow of phosphorylation cascades. In this section, we will describe how phosphorylation can be implicated in this type of MAPK ERK1/2 signaling regulation. Scaffold proteins bind to multiple interacting proteins by interconnecting them into a stable complex. This allows the rapid transmission of the signal. Another role of scaffolds is to sequester sets of interacting proteins to limit interactions with other proteins and minimize crosstalk between pathways that some components may share. Scaffold proteins such as KSR1, β-Arrestin, paxillin and IQGAP1 regulate the kinetics, amplitude, and localization of ERK1/2 signaling [30]. Ras-1 suppressor kinase (KSR1) is one of the best characterized scaffold proteins in the ERK1/2 cascade. It has several different domains through which it can interact with C-Raf, MEK1/2, and ERK1/2. In response to growth factors, KSR1 translocates to the plasma membrane where it promotes the activation of MEK1/2 by presenting it to activated Raf. In the absence of stimulus, the ubiquitin-protein isopeptide ligase family member IMP and the 14.3.3 protein prevent the function of KSR1. Mitogens induce the dephosphorylation of IMP at S392 by protein phosphatase-2A (PP2A) and the degradation of the protein, which is sufficient to allow KSR1 to translocate to the cell membrane [31]. Activated Ras also induces phosphorylation of KSR1 at residues Thr260, Thr274, and Ser443 [16]. Then, while activated Ras prevents the effects of 14.3.3 and IMP that inhibit KSR1 function, it also induces its phosphorylation at Thr274, preparing KSR1 for degradation. KSR1 can then regulate ERK1/2 activation kinetics and influence the biological fate of the cell. The interaction and in particular the synchronization between these processes generates a combinatorial control to modulate both the amplitude and the duration of ERK1/2 activity.

If scaffold proteins play a key role in regulating ERK1/2 signaling in subcellular locations, different factors modulate the strength and the duration of ERK signaling in time: the density of cell surface receptor and its different internalization patterns, the surrounding extracellular matrix and the interaction between kinases and phosphatases. The duration of the signal is critical in determining cell response to ERK1/2 signaling. For instance, long-term ERK1/2 activation can cause differentiation while short-term ERK1/2 activation can lead to cell division. This was initially demonstrated in rat pheochromocytoma PC-12 cells, in which transient activation of ERK1/2 by epidermal growth factor (EGF) or insulin peaks at 5 min and fells back to near-background levels within 15 minutes, and results in cell proliferation. On the other hand, sustained activation of ERK1/2 by nerve growth factor (NGF) persists for more than 60 minutes and induces cell differentiation [32]. This type of cell response according to duration of ERK1/2 signaling has been also reported in fibroblasts, macrophages and T lymphocytes [33, 34, 35]. As this type of studies has been made using mainly immunoblotting techniques to monitor ERK1/2 activation dynamics, the use of new approaches gaining spatio-temporal resolution will be of great interest to advance in the understanding of ERK1/2 signaling in time and in subcellular localizations. For example, using Förster Resonance Energy Transfer (FRET)-based ERK biosensors, Keyes et al. showed that EGF induces sustained ERK1/2 activity near the plasma membrane in contrast to the transient activity observed in the cytoplasm and in the nucleus. This supports the concept that the spatial and temporal regulation of ERK1/2 activity is integrated by the cell to control the specificity of signaling [36].

Studies on RTK receptors have shown that their activation kinetics and regulatory mechanisms also play a key role in the activation of the MAPK ERK1/2 pathway. For example, PC-12 cells that express few NGF receptors do not undergo differentiation in response to NGF [37]. Moreover, changing the amount of receptor occupation by decreasing the concentration of agonists alters the duration of ERK1/2 signaling. The rate and degree of receptor internalization also contribute to ERK signaling, not only as a checkpoint for signal termination, but may exhibit additional signaling by the receptor-ligand complex from an internalized cellular location [38].

Advertisement

5. Dysregulation of MAPK ERK1/2 signaling in human cancer

The MAPK/ERK signaling module is considered the most important oncogenic driver of human malignancies [39]. Mutational oncogenic activation of the Ras/Raf/MEK/ERK pathway occurs in a wide variety of cancers concerning approximately 34% of all human cancers. Activation of the ERK1/2 signaling pathway promotes proliferation and has anti-apoptotic effects, increasing tumor invasion and metastasis. The overexpression of the pathway can lead to cell transformation, tumor proliferation, invasion, metastasis, extracellular matrix degradation and tumor angiogenesis. VEGF is an important pro-angiogenic factor and the most powerful pro-vascular endothelial growth cytokine that promotes cell division and vascular construction. The MAPK ERK1/2 signaling pathway can activate transcription factors to enhance the transcription of VEGF, promoting the formation of blood vessels and tumor angiogenesis [40, 41].

Aberrant activation of the Ras/Raf/MEK/ERK pathway may be driven by abnormal receptor kinase activation or by oncogenic mutations of pathway components, leading to tumorigenesis. Overactivation of Ras is observed in approximately 30% of all human cancers but can be higher in some cancers like pancreas cancer (90%), colon cancer (50%) and thyroid cancer (50%) [42, 43]. Mutations in Ras occurs in codons 12, 13, 59 and 61, leading to its constitutive activation. Indeed, mutant oncogenic Ras proteins are insensitive to GTP-catalyzed GTPase hydrolysis activator protein, resulting in a constitutively active GTP-bound Ras. K-Ras and N-Ras are the most common mutated isoforms in human cancer, although H-Ras can also be involved. K-Ras is involved in up to 96% of pancreatic ductal adenocarcinomas, 52% of colorectal carcinomas and 32% of lung adenocarcinomas [44].

Downstream of Ras, Raf can be activated by mutations that mainly affect B-Raf isoform, the most potent activator of MEK1/2 compared with the other Raf isoforms (A-Raf and C-Raf). B-Raf can be mutated in 70% of melanomas, in 36-53% of papillary thyroid cancer, in 30% of ovarian cancer and in 22% of colorectal cancer [45]. The most common mutation of B-Raf is the change of a valine to a glutamic acid in position 600 (V600E). Other B-Raf mutations in cancer are mainly clustered in the activation segment or the so-called glycine-rich loop in B-Raf [46]. Oncogenic mutations of B-Raf lead to hyperactivity of its downstream effectors MEK1/2 and ERK1/2. For cellular transformation to occur, two mutations in Ras/Raf/MEK/ERK1/2 pathway can be needed: for instance, B-Raf and Ras mutations can drive tumorigenesis for colorectal cancer (K-ras G13D; B-Raf G463V), for ovarian cancer (K-ras G13D; B-Raf G463E), and for non-small cell lung cancer (N-Ras Q61K; B-Raf L596V) [45].

Downstream of Raf, MEK1/2 can be highly phosphorylated in colorectal cancer, gliomas, prostate cancer, breast cancer and head and neck cancer [47, 48, 49, 50, 51]. Constitutively active mutants of MEK-1 have higher basal activity than the wildtype unphosphorylated MEK. Expression of these mutants in mammalian cells lead to ERK1/2 activation in growth factor-deprived cells, cellular transformation and solid tumor growth in nude mice [33, 52, 53]. If mutant MEK can act as oncogene, its frequency in human cancers appears to be rare [54]. Finally, MAPK ERK1/2 are not frequently mutated. However, some mutations in ERK have been described: ERK2 mutants were identified as rare cancer-associated gain- and loss-of-function gene products: ERK2 D321N, ERK2 E322K, ERK2 L73P, ERK2 S151D and ERK2 D319N [55, 56, 57, 58, 59, 60]. While ERK2 D319N has not an increased basal kinase activity, it shows an elevated sensitivity to low levels of signaling in vivo [55]. In human cancer cell lines, ERK2 E322K has constitutive phosphorylation [61]. Finally, ERK2 L73P and S151D mutations increase by 8-to-12-fold ERK2 activity alone, and both mutations have a synergetic action that increases by 50-fold ERK2 activity [57]. Moreover, overexpression of ERK2-L73P/S151D can induce growth arrest in prostate cancer cell lines [62]. Although ERK1/2 mutations are rare, mutations that lead to overactivation of RTK, Ras, Raf and MEK can lead to increased ERK1/2 signaling in cancer cells. Downstream of ERK1/2, both cytoplasmic and nuclear targets can be upregulated in tumoral contexts. One of the main cancer-associated ERK substrates is c-Myc, a transcriptional factor that participates in cell cycle progression, becoming an oncogene. Phosphorylation of c-Myc by ERK1/2 due to Ras activation keeps overexpressed this transcriptional factor in various cancers [63]. Other important targets of ERK are the transcriptional factor Elk1, c-Fos and Jun. These two latter were identified as viral oncoproteins and can play a role in tumorigenesis. Mutations that affect MAPK ERK1/2 proteins can then promote protein hyperactivation that induces the cascade of phosphorylation downstream events, favoring cell proliferation, cell transformation and the emergence and progression of tumors. Currently, MAPK inhibitors represent specific target treatments for cancers with overactivation of this cell signaling pathway.

Advertisement

6. MAPK ERK1/2 inhibitors: possibility to regulate cell signaling overactivation

Hyperactivation of Ras/Raf/MEK/ERK signaling pathway in human cancers prompted the development of small molecule inhibitors that target its components for use in cancer therapeutics (Table 2). Pharmacological inhibition of Ras has been a major challenge. For instance, the affinity of Ras protein for GTP is extraordinarily high and it is then very difficult to develop a competitive binding strategy. Over the past few years, several groups discovered and developed small molecule Ras modulators using protein structure-guided design approaches [80, 81, 82] and exploring SOS as a target for Ras activation [83]. Cysteine-reactive inhibitors that bind to the mutant K-Ras G12C, which is commonly found in cancer, have been developed: SML-8-73-1 and SML-10-70-1 can selectively inhibit K-Ras G12C, changing the nucleotide preference to favor GDP over GTP and thus blocking Ras signaling [69, 84]. These compounds may be used in the future for additional K-Ras mutations.

ProteinMutationCancerInhibitorTest/effect/approvalReference
B-RafV600E V600K V226MMelanoma (66%) Ovarian cancer (35-70%) Thyroid cancer (70%)VemurafenibApproved by the FDA for metastatic and unresectable melanoma with B-Raf V600K mutationFDA
Dabrafenib and TrametinibApproved by the FDA and EMA for melanoma cancer, anaplastic thyroid cancer, NSCLCFDA, EMA [64]
LGX818Approved by the FDA for the treatment of patients with unresectable or metastatic melanoma with B-Raf mutationsFDA [65]
TAK-632TAK-632 demonstrates potent antiproliferative effects both on NRAS-mutated melanoma cells and B-Raf-mutated melanoma cells; the combination of TAK-632 and the MAPK kinase (MEK) inhibitor TAK-733 exhibits synergistic antiproliferative effects on these cells[66]
MLN2480In vitro analysis of MLN2480 and TAK-733 (allosteric MEK kinase inhibitor) demonstrates synergistic activity in cell proliferation. In vivo, MLN2480 shows antitumor activity in melanoma, colon, lung, and pancreatic cancer xenograft models[67, 68]
SorafenibApproved by the FDA for renal and hepatocellular carcinomaFDA
N-RasQ61RQ61LG12DMelanoma (15- 20%) Myeloid leukemia (30%) Lung cancer (35%)Ribociclib and BinimetinibPhase Ib/II trials in patients with locally advanced or metastatic N-Ras mutant melanomaClinical trial NCT01781572
G12VThyroid carcinoma (27%)Trametinib and PalbociclibPhase I/II trial in patients with solid tumors and with a specific cohort for N-Ras-mutant melanomaClinical trial NCT02065063
K-RasG12DG12CG13DLung cancer (30%) Colorectal cancer (45%) Pancreatic cancer (90%) Blandder cancer (50%)SML-8-73-1, SML-10-70-1SML-10-70-1, a prodrug of SML-8-73-1, inhibits lung cancer A549, H23, and H358 cells[69]
MEKS217ES221EMelanoma (3-8%)
Breast cancer (7-9%)
SelumetinibApproved by the FDA for treatment of pediatric patients aged 2 years and older with neurofibromatosis type 1 (NF1) who have symptomatic, inoperable plexiform neurofibromas; approved by the EMA for the treatment of neurofibromatosisFDA, EMA
TrametinibApproved by the FDA and EMA for treatment of patients with unresectable or metastatic melanoma with B-Raf V600E or V600K mutationsFDA, EMA
PimasertibHas demonstrated potent antitumour activity in human lung, colorectal, melanoma cancer cells and xenograft models; phase I/II clinical trial in patients with locally advanced or metastatic solid tumors[70, 71, 72]
CobimetinibApproved by the FDA and EMA for use in combination with vemurafenib for the treatment of metastatic melanomaFDA, EMA
G-573, GDC-0623In vitro GDC-0623 inhibits cellular proliferation of mutant cancer cells A375 (B-Raf V600E), HCT116 (KRAS G13D), COLO 205 (BRAF V600E), HT- 29 (BRAF V600E), and HCT116 (KRAS G13D). In vivo GDC-0623 causes potent tumor growth inhibition in mouse MiaPaCa-2, A375 and HCT116 xenografts[73]
TAK-733In vitro TAK-733 demonstrates broad activity in most melanoma cell lines; in vivo TAK-733 demonstrates broad antitumor activity in mouse xenograft models of human cancer including melanoma, colorectal, NSCLC, pancreatic and breast cancer[74]
Binimetinib(Mek 162)Approved by the FDA for treatment in combination with LGX818 for patients with unresectable or metastatic melanoma with a B-Raf V600E or V600K mutation; approved by the EMA for the treatment of colorectal carcinomaFDA, EMA
ERK 1/2NAMelanoma (67%)Ulixertinib(BVD-523)In vitro combined Ulixertinib (BVD-523) and VS-5584 treatment causes significant induction of cell death in human pancreatic cancer (HPAC) cells, in pancreatic ductal adenocarcinoma cell lines BxPC-3, MIAPaCa-2, and CFPAC-1. Clinical trials in phase I for tumor advanced pancreatic and other solid tumors cancer and phase II for advanced malignancies harboring MEK or atypical B-Raf alterations.[75] Clinical trial NCT03454035 and NCT04488003
GDC-0994In vitro Ravoxertinib (GDC0994) decreases the viability of lung adenocarcinoma cell lines (A549, HCC827, HCC4006). In vivo GDC0994 results in significant single-agent activity in multiple cancer models, including K-Ras-mutant and B-Raf-mutant human xenograft tumors in mice. Clinical trials phase I for locally advanced or metastatic solid tumors, NSCLC, metastatic colorectal cancer, metastatic NSCLC, metastatic cancers and melanoma[76] Clinical trials NCT01875705 and NCT02457793
SCH772984In vitro SCH772984 results in a G1 arrest in SCH772984-sensitive melanoma cells. In vivo antitumor activity is observed in the K-Ras-mutant pancreatic MiaPaCa model[77]
AEZS-134Synergistic effect of triptorelin, ERK inhibitor AEZS-134 and dual PI3K/ERK inhibitor AEZS-136 in MDA-MB-231 triple-negative breast cancer cells[78]
(S)-14 KIn vivo (S)-14 k inhibited tumor growth in mouse xenograft models[79]

Table 2.

Available inhibitors for MAPK ERK1/2 proteins.

NA: not available; FDA: food and drug administration of USA; EMA: European medicines agency; NSCLC: non-small cell lung cancer.

Sorafenib is an orally available compound that was initially developed as a C-Raf inhibitor and was then identified as a multikinase inhibitor for B-Raf, VEGFR1/2/3, Kit, PDGFR, RET, and Flt3. Sorafenib is currently approved by the FDA for renal and hepatocellular carcinoma for its anti-angiogenic effects [84, 85]. For other cancers like melanoma, sorafenib produced favorable responses in less than 5% of patients in clinical trials [85, 86]. This low response rate can be due to the fact that its activity against B-Raf V600E mutants and wild-type enzymes is low. Subsequent efforts have focused on targeting B-Raf for the treatment of B-Raf mutant melanoma. Vemurafenib and dabrafenib, two B-Raf V600E inhibitors, have achieved benefits in clinical trials [87, 88]. Currently, vemurafenib is approved by the FDA for metastatic and unresectable melanoma with B-Raf V600K mutation [89] and dabrafenib for metastatic melanoma with B-Raf V600K-mutated [84, 88]. Although B-Raf inhibitors have achieved clinical benefit in the treatment of cancer, all ATP-competitive Raf inhibitors including vemurafenib, dabrabenib, and sorafenib can lead to paradoxical activation of the MAPK pathway in wild-type B-Raf cells [90, 91]. Some reports suggest that insensitivity to Raf inhibitors might be due to EGFR-mediated reactivation of MAPK signaling in B-Raf mutant colorectal cancer [92]. Indeed, the combination of EGFR and B-Raf inhibitors block the reactivation of MAPK signaling of B-Raf mutant in colorectal cancer cells and in vivo [93]. LGX818, TAK-632 and MLN2480 are other selective B-Raf V600E inhibitors with a very slow inactivation rate, and thus may be beneficial for the treatment of tumors that are resistant to other Raf inhibitors or for the treatment of tumors with Ras mutations [66, 94].

Even though MEK1/2 mutations are rare in human cancers, MEK1/2 have become an attractive drug target because these proteins are downstream of Ras and Raf in the signaling pathway [95]. The first MEK1/2 inhibitor, PD098059, is an allosteric inhibitor that acts on the not-phosphorylated form of MEK1 and mutant MEK1 S217 and S221E [96]. The allosteric MEK inhibitor CI-1040 was the first small molecule to enter clinical trials. Although it had antitumor effects, the development of this compound was stopped due to poor bioavailability and lack of efficacy in phase II clinical trials [97]. Other highly selective inhibitors of MEK1 and MEK2 include selumetinib and trametinib [98, 99, 100]. This latter prevents Raf-dependent MEK phosphorylation and activation. Other MEK inhibitors in development include pimasertib [101], cobimetinib [102], rafametinib [103], G-573, GDC-0623 [73], TAK-733 [104], RO5126766, RO4987655 [105, 106] and MEK162 [107].

Because there are few ERK1/2 mutations in human cancers, this MAPK has been only considered as a target in 35 clinical trials, compared with more than 300 clinical trials for the inhibition of Raf and MEK. Nonetheless, due to drug resistance resulting from Raf and MEK1/2 inhibitors, ERK1/2 have become an interesting target for inhibiting MAPK ERK1/2 signaling in cancer [46]. ERK1/2 inhibitors can reverse overactivation of the MAPK pathway induced by upstream mutations, including Ras mutations [84, 92, 108]. For instance, MAPK inhibition in B-Raf V600E mutant metastatic melanoma provokes drug resistance and recovery of ERK activity [109, 110]. Interestingly, selective removal of ERK1 or ERK2 in vitro can induce melanoma cell death and enhances the action of B-Raf inhibitor [111].

One of the challenges in cancer treatment is developing drug resistance. The mechanisms involved in resistance are complicated and include genetic mutations that occur in target proteins like in MAPK signaling, loss of functions in the control of MAPK signaling feedback, and abnormal tumor suppressor gene alterations [112]. Yet, MAPK inhibitors represent good options for targeting cancer cells with MAPK overactivation or MAPK ERK1/2 mutations. In the future, cell-specific deliverance of MAPK inhibitors to tumoral cells should enhance their efficiency and decrease side effects in patients.

Advertisement

7. Conclusions

MAPK are conserved kinases in eukaryotes, containing 3-tier kinases that are sequentially activated by phosphorylation. This post-translational modification plays an essential role in MAPK ERK1/2 signaling. Not only the activation but also the regulation of this pathway is achieved through the actions of kinases and phosphatases, establishing positive and negative signaling feedbacks. Control of MAPK ERK1/2 signaling in time and space is ensured by proteins such as scaffolds that are themselves regulated by phosphorylation events. Changes in duration of ERK1/2 phosphorylation and thus activity, can result in different cell responses, can result in different cell responses. Thus, a tight regulation of MAPK ERK1/2 signaling is needed to guarantee adaptive cell responses. Aberrant activation of Ras/Raf/MEK/ERK pathway can lead to tumorigenesis and MAPK inhibitors, already in clinical use, represent good options for targeting cancer cells with MAPK overactivation or MAPK ERK1/2 mutations.

Advertisement

Acknowledgments

We thank Helden Natalia Vélez González for assistance in figure design.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Appendices and nomenclature

DUSPdual specificity phosphatase
EGFepidermal growth factor
EMAEuropean medicines agency
ERKextracellular signal-regulated kinases
FDAfood and drug administration
FRETFörster resonance energy transfer
GRB2growth factor receptor-bound protein 2
KSR1Ras-1 suppressor kinase
MAPKmitogen-protein activated kinases
MAPKKMAPK kinase
MAPKKKMAPK kinase kinase
MKPmap kinase phosphatase
mTORmechanistic target of rapamycin
NGFnerve growth factor
NSCLCnon-small cell lung cancer
PDGF-Bplatelet-derived growth factor subunit B
PDGFRplatelet-derived growth factor
PI3Kphosphatidylinositol 4,5-bisphosphate 3-kinase
PPMmetal-dependent protein phosphatase
PPPphosphoprotein phosphatase
PRSproline-rich sequence
PTENphosphatidylinositol 3,4,5-trisphosphate 3-phosphatase and dual-specificity protein phosphatase
PTPprotein tyrosine phosphatases
RTKreceptor tyrosine kinase
RSK-2ribosomal S6 kinase 2
SOS1son of sevenless homolog 1

References

  1. 1. Levene PA, Alsberg CL. The Cleavage Products of Vitellin. J Biol Chem. 1906 Aug 1;2(1):127-133
  2. 2. Lipmann FA, Levene PA. Serinephosphoric acid obtained on hydrolysis of vitellinic acid. J Biol Chem. 1932 Oct 1;98(1):109-114
  3. 3. Burnett G, Kennedy EP. The enzymatic phosphorylation of proteins. J Biol Chem. 1954 Dec;211(2):969-980
  4. 4. Fischer EH, Krebs EG. Conversion of phosphorylase b to phosphorylase a in muscle extracts. J Biol Chem. 1955 Sep;216(1):121-132
  5. 5. Chen MJ, Dixon JE, Manning G. Genomics and evolution of protein phosphatases. Sci Signal [Internet]. 2017 Apr 11 [cited 2021 Feb 7];10(474). Available from: https://stke.sciencemag.org/content/10/474/eaag1796
  6. 6. Lavoie H, Gagnon J, Therrien M. ERK signalling: a master regulator of cell behaviour, life and fate. Nat Rev Mol Cell Biol. 2020 Oct;21(10):607-632
  7. 7. Jacobs D, Glossip D, Xing H, Muslin AJ, Kornfeld K. Multiple docking sites on substrate proteins form a modular system that mediates recognition by ERK MAP kinase. Genes Dev. 1999 Jan 15;13(2):163-175
  8. 8. Ünal EB, Uhlitz F, Blüthgen N. A compendium of ERK targets. FEBS Lett. 2017 Sep;591(17):2607-2615
  9. 9. Corbalan-Garcia S, Yang SS, Degenhardt KR, Bar-Sagi D. Identification of the mitogen-activated protein kinase phosphorylation sites on human Sos1 that regulate interaction with Grb2. Mol Cell Biol. 1996 Oct;16(10):5674-5682
  10. 10. Kamioka Y, Yasuda S, Fujita Y, Aoki K, Matsuda M. Multiple Decisive Phosphorylation Sites for the Negative Feedback Regulation of SOS1 via ERK *. J Biol Chem. 2010 Oct 22;285(43):33540-33548
  11. 11. Saha M, Carriere A, Cheerathodi M, Zhang X, Lavoie G, Rush J, et al. RSK phosphorylates SOS1 creating 14-3-3-docking sites and negatively regulating MAPK activation. Biochem J. 2012 Oct 1;447(1):159-166
  12. 12. Ren J, Cook AA, Bergmeier W, Sondek J. A negative-feedback loop regulating ERK1/2 activation and mediated by RasGPR2 phosphorylation. Biochem Biophys Res Commun. 2016 May 20;474(1):193-198
  13. 13. Eblen ST, Slack-Davis JK, Tarcsafalvi A, Parsons JT, Weber MJ, Catling AD. Mitogen-activated protein kinase feedback phosphorylation regulates MEK1 complex formation and activation during cellular adhesion. Mol Cell Biol. 2004 Mar;24(6):2308-2317
  14. 14. Coles LC, Shaw PE. PAK1 primes MEK1 for phosphorylation by Raf-1 kinase during cross-cascade activation of the ERK pathway. Oncogene. 2002 Mar 28;21(14):2236-2244
  15. 15. Slack-Davis JK, Eblen ST, Zecevic M, Boerner SA, Tarcsafalvi A, Diaz HB, et al. PAK1 phosphorylation of MEK1 regulates fibronectin-stimulated MAPK activation. J Cell Biol. 2003 Jul 21;162(2):281-291
  16. 16. Cacace AM, Michaud NR, Therrien M, Mathes K, Copeland T, Rubin GM, et al. Identification of constitutive and ras-inducible phosphorylation sites of KSR: implications for 14-3-3 binding, mitogen-activated protein kinase binding, and KSR overexpression. Mol Cell Biol. 1999 Jan;19(1):229-240
  17. 17. McKay MM, Ritt DA, Morrison DK. Signaling dynamics of the KSR1 scaffold complex. Proc Natl Acad Sci. 2009 Jul 7;106(27):11022-11027
  18. 18. Brunet A, Gilles Pagès, Jacques Pouysségur. Growth factor-stimulated MAP kinase induces rapid retrophosphorylation and inhibition of MAP kinase kinase (MEK1). FEBS Lett. 1994;346(2-3):299-303
  19. 19. Rossomando AJ, Dent P, Sturgill TW, Marshak DR. Mitogen-activated protein kinase kinase 1 (MKK1) is negatively regulated by threonine phosphorylation. Mol Cell Biol. 1994 Mar;14(3):1594-1602
  20. 20. Dougherty MK, Müller J, Ritt DA, Zhou M, Zhou XZ, Copeland TD, et al. Regulation of Raf-1 by direct feedback phosphorylation. Mol Cell. 2005 Jan 21;17(2):215-224
  21. 21. Jurek A, Amagasaki K, Gembarska A, Heldin C-H, Lennartsson J. Negative and positive regulation of MAPK phosphatase 3 controls platelet-derived growth factor-induced Erk activation. J Biol Chem. 2009 Feb 13;284(7):4626-4634
  22. 22. Marchetti S, Gimond C, Chambard J-C, Touboul T, Roux D, Pouysségur J, et al. Extracellular signal-regulated kinases phosphorylate mitogen-activated protein kinase phosphatase 3/DUSP6 at serines 159 and 197, two sites critical for its proteasomal degradation. Mol Cell Biol. 2005 Jan;25(2):854-864
  23. 23. Carey KD, Watson RT, Pessin JE, Stork PJS. The Requirement of Specific Membrane Domains for Raf-1 Phosphorylation and Activation. J Biol Chem. 2003 Jan;278(5):3185-3196
  24. 24. Corbit KC, Trakul N, Eves EM, Diaz B, Marshall M, Rosner MR. Activation of Raf-1 signaling by protein kinase C through a mechanism involving Raf kinase inhibitory protein. J Biol Chem. 2003 Apr 11;278(15):13061-13068
  25. 25. Kwang-Hyun C, Sung-Young S, Hyun-Woo K, Wolkenhauer O, McFerran B, Kolch W. Mathematical Modeling of the Influence of RKIP on the ERK Signaling Pathway. In: Priami C, editor. Computational Methods in Systems Biology. Berlin, Heidelberg: Springer; 2003. p. 127-41. (Lecture Notes in Computer Science)
  26. 26. Shin S-Y, Rath O, Choo S-M, Fee F, McFerran B, Kolch W, et al. Positive- and negative-feedback regulations coordinate the dynamic behavior of the Ras-Raf-MEK-ERK signal transduction pathway. J Cell Sci. 2009 Feb 1;122(Pt 3):425-435
  27. 27. Camps M, Nichols A, Gillieron C, Antonsson B, Muda M, Chabert C, et al. Catalytic activation of the phosphatase MKP-3 by ERK2 mitogen-activated protein kinase. Science. 1998 May 22;280(5367):1262-1265
  28. 28. Muda M, Theodosiou A, Gillieron C, Smith A, Chabert C, Camps M, et al. The mitogen-activated protein kinase phosphatase-3 N-terminal noncatalytic region is responsible for tight substrate binding and enzymatic specificity. J Biol Chem. 1998 Apr 10;273(15):9323-9329
  29. 29. Bermudez O, Marchetti S, Pagès G, Gimond C. Post-translational regulation of the ERK phosphatase DUSP6/MKP3 by the mTOR pathway. Oncogene. 2008 Jun 12;27(26):3685-3691
  30. 30. Miningou N, Blackwell K. The road to ERK activation: Do neurons take alternate routes? Cell Signal. 2020 Apr 1;68:109541
  31. 31. Müller J, Ory S, Copeland T, Piwnica-Worms H, Morrison DK. C-TAK1 regulates Ras signaling by phosphorylating the MAPK scaffold, KSR1. Mol Cell. 2001 Nov;8(5):983-993
  32. 32. Traverse S, Gomez N, Paterson H, Marshall C, Cohen P. Sustained activation of the mitogen-activated protein (MAP) kinase cascade may be required for differentiation of PC12 cells. Comparison of the effects of nerve growth factor and epidermal growth factor. Biochem J. 1992 Dec 1;288(2):351-355
  33. 33. Cowley S, Paterson H, Kemp P, Marshall CJ. Activation of MAP kinase kinase is necessary and sufficient for PC12 differentiation and for transformation of NIH 3T3 cells. Cell. 1994 Jun 17;77(6):841-852
  34. 34. Sharp LL, Schwarz DA, Bott CM, Marshall CJ, Hedrick SM. The influence of the MAPK pathway on T cell lineage commitment. Immunity. 1997 Nov;7(5):609-618
  35. 35. Whalen AM, Galasinski SC, Shapiro PS, Nahreini TS, Ahn NG. Megakaryocytic differentiation induced by constitutive activation of mitogen-activated protein kinase kinase. Mol Cell Biol. 1997 Apr;17(4):1947-1958
  36. 36. Keyes J, Ganesan A, Molinar-Inglis O, Hamidzadeh A, Zhang J, Ling M, et al. Signaling diversity enabled by Rap1-regulated plasma membrane ERK with distinct temporal dynamics. eLife [Internet]. 2020 [cited 2021 Feb 14];9. Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7289600/
  37. 37. Schlessinger J, Bar-Sagi D. Activation of Ras and other signaling pathways by receptor tyrosine kinases. Cold Spring Harb Symp Quant Biol. 1994;59:173-179
  38. 38. Murphy LO, MacKeigan JP, Blenis J. A network of immediate early gene products propagates subtle differences in mitogen-activated protein kinase signal amplitude and duration. Mol Cell Biol. 2004 Jan;24(1):144-153
  39. 39. McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Wong EWT, Chang F, et al. Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim Biophys Acta BBA - Mol Cell Res. 2007 Aug 1;1773(8):1263-1284
  40. 40. Fish JE, Cantu Gutierrez M, Dang LT, Khyzha N, Chen Z, Veitch S, et al. Dynamic regulation of VEGF-inducible genes by an ERK/ERG/p300 transcriptional network. Dev Camb Engl. 2017 Jul 1;144(13):2428-2444
  41. 41. Song M, Finley SD. Mechanistic insight into activation of MAPK signaling by pro-angiogenic factors. BMC Syst Biol. 2018 Dec 27;12(1):145
  42. 42. Hobbs GA, Der CJ, Rossman KL. RAS isoforms and mutations in cancer at a glance. J Cell Sci. 2016 Apr 1;129(7):1287-1292
  43. 43. Malumbres M, Barbacid M. RAS oncogenes: the first 30 years. Nat Rev Cancer. 2003 Jun;3(6):459-465
  44. 44. Drosten M, Barbacid M. Targeting the MAPK Pathway in KRAS-Driven Tumors. Cancer Cell. 2020 Apr 13;37(4):543-550
  45. 45. Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, et al. Mutations of the BRAF gene in human cancer. Nature. 2002 Jun;417(6892):949-954
  46. 46. Samatar AA, Poulikakos PI. Targeting RAS-ERK signalling in cancer: promises and challenges. Nat Rev Drug Discov. 2014 Dec;13(12):928-942
  47. 47. Gee JM, Robertson JF, Ellis IO, Nicholson RI. Phosphorylation of ERK1/2 mitogen-activated protein kinase is associated with poor response to anti-hormonal therapy and decreased patient survival in clinical breast cancer. Int J Cancer. 2001 Jul 20;95(4):247-254
  48. 48. Gioeli D, Mandell JW, Petroni GR, Frierson HF, Weber MJ. Activation of mitogen-activated protein kinase associated with prostate cancer progression. Cancer Res. 1999 Jan 15;59(2):279-284
  49. 49. Gomez-Millan J, Pajares B, Perez-Villa L, Carnero A, Alvarez M, De Luque V, et al. Subcellular localisation of pMEK has a different prognosis in locally advanced head and neck cancer treated with concomitant radiochemotherapy. BMC Cancer. 2016 Oct 28;16(1):829
  50. 50. Lee SH, Lee JW, Soung YH, Kim SY, Nam SW, Park WS, et al. Colorectal tumors frequently express phosphorylated mitogen-activated protein kinase. APMIS. 2004;112(4-5):233-238
  51. 51. Mandell JW, Hussaini IM, Zecevic M, Weber MJ, VandenBerg SR. In situ visualization of intratumor growth factor signaling: immunohistochemical localization of activated ERK/MAP kinase in glial neoplasms. Am J Pathol. 1998 Nov;153(5):1411-1423
  52. 52. Brunet A, Pagès G, Pouysségur J. Constitutively active mutants of MAP kinase kinase (MEK1) induce growth factor-relaxation and oncogenicity when expressed in fibroblasts. Oncogene. 1994 Nov;9(11):3379-3387
  53. 53. Mansour SJ, Matten WT, Hermann AS, Candia JM, Rong S, Fukasawa K, et al. Transformation of mammalian cells by constitutively active MAP kinase kinase. Science. 1994 Aug 12;265(5174):966-970
  54. 54. Bansal A, Ramirez RD, Minna JD. Mutation analysis of the coding sequences of MEK-1 and MEK-2 genes in human lung cancer cell lines. Oncogene. 1997 Mar 13;14(10):1231-1234
  55. 55. Bott CM, Thorneycroft SG, Marshall CJ. The sevenmaker gain-of-function mutation in p42 MAP kinase leads to enhanced signalling and reduced sensitivity to dual specificity phosphatase action. FEBS Lett. 1994 Sep 26;352(2):201-205
  56. 56. Brenan L, Andreev A, Cohen O, Pantel S, Kamburov A, Cacchiarelli D, et al. Phenotypic Characterization of a Comprehensive Set of MAPK1/ERK2 Missense Mutants. Cell Rep. 2016 Oct 18;17(4):1171-1183
  57. 57. Emrick MA, Hoofnagle AN, Miller AS, Ten Eyck LF, Ahn NG. Constitutive activation of extracellular signal-regulated kinase 2 by synergistic point mutations. J Biol Chem. 2001 Dec 7;276(49):46469-46479
  58. 58. Goetz EM, Ghandi M, Treacy DJ, Wagle N, Garraway LA. ERK mutations confer resistance to mitogen-activated protein kinase pathway inhibitors. Cancer Res. 2014 Dec 1;74(23):7079-7089
  59. 59. Lawrence MS, Stojanov P, Mermel CH, Robinson JT, Garraway LA, Golub TR, et al. Discovery and saturation analysis of cancer genes across 21 tumour types. Nature. 2014 Jan 23;505(7484):495-501
  60. 60. Ojesina AI, Lichtenstein L, Freeman SS, Pedamallu CS, Imaz-Rosshandler I, Pugh TJ, et al. Landscape of genomic alterations in cervical carcinomas. Nature. 2014 Feb 20;506(7488):371-375
  61. 61. Arvind R, Shimamoto H, Momose F, Amagasa T, Omura K, Tsuchida N. A mutation in the common docking domain of ERK2 in a human cancer cell line, which was associated with its constitutive phosphorylation. Int J Oncol. 2005 Dec;27(6):1499-1504
  62. 62. Wu P-K, Hong S-K, Yoon S-H, Park J-I. Active ERK2 is sufficient to mediate growth arrest and differentiation signaling. FEBS J. 2015;282(6):1017-1030
  63. 63. Sears R, Nuckolls F, Haura E, Taya Y, Tamai K, Nevins JR. Multiple Ras-dependent phosphorylation pathways regulate Myc protein stability. Genes Dev. 2000 Oct 1;14(19):2501-2514
  64. 64. Odogwu L, Mathieu L, Blumenthal G, Larkins E, Goldberg KB, Griffin N, et al. FDA Approval Summary: Dabrafenib and Trametinib for the Treatment of Metastatic Non-Small Cell Lung Cancers Harboring BRAF V600E Mutations. The Oncologist. 2018;23(6):740-745
  65. 65. Li Z, Jiang K, Zhu X, Lin G, Song F, Zhao Y, et al. Encorafenib (LGX818), a potent BRAF inhibitor, induces senescence accompanied by autophagy in BRAFV600E melanoma cells. Cancer Lett. 2016 Jan 28;370(2):332-344
  66. 66. Nakamura A, Arita T, Tsuchiya S, Donelan J, Chouitar J, Carideo E, et al. Antitumor activity of the selective pan-RAF inhibitor TAK-632 in BRAF inhibitor-resistant melanoma. Cancer Res. 2013 Dec 1;73(23):7043-7055
  67. 67. Cunniff EGC, Zhang J, Chouitar J, Mettetal J, Nakamura K, Arita T, et al. Abstract C146: Combination treatment with the investigational RAF kinase inhibitor MLN2480 and the investigational MEK kinase inhibitor TAK-733 inhibits the growth of BRAF mutant and RAS mutant preclinical models of melanoma and CRC. Mol Cancer Ther. 2013 Nov 1;12(11 Supplement):C146–C146
  68. 68. Rasco DW, Olszanski AJ, Patnaik A, Espino G, Neuwirth R,Faucette S, et al. MLN2480, an investigational oral pan-RAF kinase inhibitor, in patients (pts) with relapsed or refractory solid tumors: Phase I study. J Clin Oncol. 2013 May 20;31(15_suppl):2547-2547
  69. 69. Lim SM, Westover KD, Ficarro SB, Harrison RA, Choi HG, Pacold ME, et al. Therapeutic Targeting of Oncogenic K-Ras by a Covalent Catalytic Site Inhibitor. Angew Chem. 2014;126(1):203-208
  70. 70. Martinelli E, Troiani T, D’Aiuto E, Morgillo F, Vitagliano D, Capasso A, et al. Antitumor activity of pimasertib, a selective MEK 1/2 inhibitor, in combination with PI3K/mTOR inhibitors or with multi-targeted kinase inhibitors in pimasertib-resistant human lung and colorectal cancer cells. Int J Cancer. 2013;133(9):2089-2101
  71. 71. Delord JP, Houédé N, Awada A, Lebbe C, Lesimple T, Schellens JHM, et al. 616 Pimasertib (MSC1936369B/AS703026), a Selective Oral MEK1/2 Inhibitor, Shows Clinical Activity in Melanoma. Eur J Cancer. 2012 Nov 1;48:190
  72. 72. Infante JR, Gandhi L, Shapiro G, Burris HA, Bendell JC, Baselga J, et al. Phase lb combination trial of a MEK inhibitor, pimasertib (MSC1936369B), and a PI3K/mTOR inhibitor, SAR245409, in patients with locally advanced or metastatic solid tumors. J Clin Oncol. 2012 May 20;30(15_suppl):TPS3118–TPS3118
  73. 73. Hatzivassiliou G, Haling JR, Chen H, Song K, Price S, Heald R, et al. Mechanism of MEK inhibition determines efficacy in mutant KRAS- versus BRAF-driven cancers. Nature. 2013 Sep 12;501(7466):232-236
  74. 74. Micel LN, Tentler JJ, Tan A-C, Selby HM, Brunkow KL, Robertson KM, et al. Antitumor Activity of the MEK Inhibitor TAK-733 against Melanoma Cell Lines and Patient-Derived Tumor Explants. Mol Cancer Ther. 2015 Feb 1;14(2):317-325
  75. 75. Ning C, Liang M, Liu S, Wang G, Edwards H, Xia Y, et al. Targeting ERK enhances the cytotoxic effect of the novel PI3K and mTOR dual inhibitor VS-5584 in preclinical models of pancreatic cancer. Oncotarget. 2017 May 15;8(27):44295-44311
  76. 76. Robarge K, Schwarz J, Blake J, Burkard M, Chan J, Chen H, et al. Abstract DDT02-03: Discovery of GDC-0994, a potent and selective ERK1/2 inhibitor in early clinical development. Cancer Res. 2014 Oct 1;74(19 Supplement):DDT02-03-DDT02-03
  77. 77. Morris EJ, Jha S, Restaino CR, Dayananth P, Zhu H, Cooper A, et al. Discovery of a novel ERK inhibitor with activity in models of acquired resistance to BRAF and MEK inhibitors. Cancer Discov. 2013 Jul;3(7):742-750
  78. 78. Kwok CW, Treeck O, Buchholz S, Seitz S, Ortmann O, Engel JB. Receptors for luteinizing hormone-releasing hormone (GnRH) as therapeutic targets in triple negative breast cancers (TNBC). Target Oncol. 2015 Sep;10(3):365-373
  79. 79. Ren L, Grina J, Moreno D, Blake JF, Gaudino JJ, Garrey R, et al. Discovery of Highly Potent, Selective, and Efficacious Small Molecule Inhibitors of ERK1/2. J Med Chem. 2015 Feb 26;58(4):1976-1991
  80. 80. Burns MC, Sun Q, Daniels RN, Camper D, Kennedy JP, Phan J, et al. Approach for targeting Ras with small molecules that activate SOS-mediated nucleotide exchange. Proc Natl Acad Sci U S A. 2014 Mar 4;111(9):3401-3406
  81. 81. Maurer T, Garrenton LS, Oh A, Pitts K, Anderson DJ, Skelton NJ, et al. Small-molecule ligands bind to a distinct pocket in Ras and inhibit SOS-mediated nucleotide exchange activity. Proc Natl Acad Sci U S A. 2012 Apr 3;109(14):5299-5304
  82. 82. Sun Q, Burke JP, Phan J, Burns MC, Olejniczak ET, Waterson AG, et al. Discovery of small molecules that bind to K-Ras and inhibit Sos-mediated activation. Angew Chem Int Ed Engl. 2012 Jun 18;51(25):6140-6143
  83. 83. Lu S, Jang H, Zhang J, Nussinov R. Inhibitors of Ras–SOS Interactions. ChemMedChem. 2016;11(8):814-821
  84. 84. Liu F, Yang X, Geng M, Huang M. Targeting ERK, an Achilles’ Heel of the MAPK pathway, in cancer therapy. Acta Pharm Sin B. 2018 Jul 1;8(4):552-562
  85. 85. Roskoski R. Targeting ERK1/2 protein-serine/threonine kinases in human cancers. Pharmacol Res. 2019 Apr 1;142:151-168
  86. 86. Eisen T, Ahmad T, Flaherty KT, Gore M, Kaye S, Marais R, et al. Sorafenib in advanced melanoma: a Phase II randomised discontinuation trial analysis. Br J Cancer. 2006 Sep;95(5):581-586
  87. 87. Chapman PB, Hauschild A, Robert C, Haanen JB, Ascierto P, Larkin J, et al. Improved Survival with Vemurafenib in Melanoma with BRAF V600E Mutation. N Engl J Med. 2011 Jun 30;364(26):2507-2516
  88. 88. Flaherty KT, Puzanov I, Kim KB, Ribas A, McArthur GA, Sosman JA, et al. Inhibition of Mutated, Activated BRAF in Metastatic Melanoma. N Engl J Med. 2010 Aug 26;363(9):809-819
  89. 89. Bollag G, Tsai J, Zhang J, Zhang C, Ibrahim P, Nolop K, et al. Vemurafenib: the first drug approved for BRAF-mutant cancer. Nat Rev Drug Discov. 2012 Nov;11(11):873-886
  90. 90. Hatzivassiliou G, Song K, Yen I, Brandhuber BJ, Anderson DJ, Alvarado R, et al. RAF inhibitors prime wild-type RAF to activate the MAPK pathway and enhance growth. Nature. 2010 Mar 18;464(7287):431-435
  91. 91. Poulikakos P, C Z, G B, Km S, N R. RAF inhibitors transactivate RAF dimers and ERK signalling in cells with wild-type BRAF [Internet]. Vol. 464, Nature. Nature; 2010 [cited 2021 Feb 25]. Available from: https://aplicacionesbiblioteca.udea.edu.co:2598/20179705/
  92. 92. Ahronian LG, Sennott EM, Allen EMV, Wagle N, Kwak EL, Faris JE, et al. Clinical Acquired Resistance to RAF Inhibitor Combinations in BRAF-Mutant Colorectal Cancer through MAPK Pathway Alterations. Cancer Discov. 2015 Apr 1;5(4):358-367
  93. 93. Corcoran RB, Ebi H, Turke AB, Coffee EM, Nishino M, Cogdill AP, et al. EGFR-mediated re-activation of MAPK signaling contributes to insensitivity of BRAF mutant colorectal cancers to RAF inhibition with vemurafenib. Cancer Discov. 2012 Mar;2(3):227-235
  94. 94. Stuart DD, Li N, Poon DJ, Aardalen K, Kaufman S, Merritt H, et al. Abstract 3790: Preclinical profile of LGX818: A potent and selective RAF kinase inhibitor. Cancer Res. 2012 Apr 15;72(8 Supplement):3790-3790
  95. 95. Carlino MS, Todd JR, Gowrishankar K, Mijatov B, Pupo GM, Fung C, et al. Differential activity of MEK and ERK inhibitors in BRAF inhibitor resistant melanoma. Mol Oncol. 2014 May 1;8(3):544-554
  96. 96. Alessi DR, Cuenda A, Cohen P, Dudley DT, Saltiel AR. PD 098059 Is a Specific Inhibitor of the Activation of Mitogen-activated Protein Kinase Kinase in Vitro and in Vivo(*). J Biol Chem. 1995 Nov 17;270(46):27489-27494
  97. 97. Lorusso PM, Adjei AA, Varterasian M, Gadgeel S, Reid J, Mitchell DY, et al. Phase I and pharmacodynamic study of the oral MEK inhibitor CI-1040 in patients with advanced malignancies. J Clin Oncol Off J Am Soc Clin Oncol. 2005 Aug 10;23(23):5281-5293
  98. 98. Davies BR, Logie A, McKay JS, Martin P, Steele S, Jenkins R, et al. AZD6244 (ARRY-142886), a potent inhibitor of mitogen-activated protein kinase/extracellular signal-regulated kinase kinase 1/2 kinases: mechanism of action in vivo, pharmacokinetic/pharmacodynamic relationship, and potential for combination in preclinical models. Mol Cancer Ther. 2007 Aug;6(8):2209-2219
  99. 99. Lugowska I, Koseła-Paterczyk H, Kozak K, Rutkowski P. Trametinib: a MEK inhibitor for management of metastatic melanoma. OncoTargets Ther. 2015 Aug 25;8:2251-2259
  100. 100. Yeh TC, Marsh V, Bernat BA, Ballard J, Colwell H, Evans RJ, et al. Biological characterization of ARRY-142886 (AZD6244), a potent, highly selective mitogen-activated protein kinase kinase 1/2 inhibitor. Clin Cancer Res Off J Am Assoc Cancer Res. 2007 Mar 1;13(5):1576-1583
  101. 101. Kim K, Kong S-Y, Fulciniti M, Li X, Song W, Nahar S, et al. Blockade of the MEK/ERK signalling cascade by AS703026, a novel selective MEK1/2 inhibitor, induces pleiotropic anti-myeloma activity in vitro and in vivo. Br J Haematol. 2010 May;149(4):537-549
  102. 102. Choo EF, Belvin M, Boggs J, Deng Y, Hoeflich KP, Ly J, et al. Preclinical disposition of GDC-0973 and prospective and retrospective analysis of human dose and efficacy predictions. Drug Metab Dispos Biol Fate Chem. 2012 May;40(5):919-927
  103. 103. Iverson C, Larson G, Lai C, Yeh L-T, Dadson C, Weingarten P, et al. RDEA119/BAY 869766: a potent, selective, allosteric inhibitor of MEK1/2 for the treatment of cancer. Cancer Res. 2009 Sep 1;69(17):6839-6847
  104. 104. Dong Q, Dougan DR, Gong X, Halkowycz P, Jin B, Kanouni T, et al. Discovery of TAK-733, a potent and selective MEK allosteric site inhibitor for the treatment of cancer. Bioorg Med Chem Lett. 2011 Mar 1;21(5):1315-1319
  105. 105. Leijen S, Middleton MR, Tresca P, Kraeber-Bodéré F, Dieras V, Scheulen ME, et al. Phase I dose-escalation study of the safety, pharmacokinetics, and pharmacodynamics of the MEK inhibitor RO4987655 (CH4987655) in patients with advanced solid tumors. Clin Cancer Res Off J Am Assoc Cancer Res. 2012 Sep 1;18(17):4794-4805
  106. 106. Rinehart J, Adjei AA, Lorusso PM, Waterhouse D, Hecht JR, Natale RB, et al. Multicenter phase II study of the oral MEK inhibitor, CI-1040, in patients with advanced non-small-cell lung, breast, colon, and pancreatic cancer. J Clin Oncol Off J Am Soc Clin Oncol. 2004 Nov 15;22(22):4456-4462
  107. 107. Ascierto PA, Schadendorf D, Berking C, Agarwala SS, van Herpen CM, Queirolo P, et al. MEK162 for patients with advanced melanoma harbouring NRAS or Val600 BRAF mutations: a non-randomised, open-label phase 2 study. Lancet Oncol. 2013 Mar;14(3):249-256
  108. 108. Hatzivassiliou G, Liu B, O’Brien C, Spoerke JM, Hoeflich KP, Haverty PM, et al. ERK Inhibition Overcomes Acquired Resistance to MEK Inhibitors. Mol Cancer Ther. 2012 May 1;11(5):1143-1154
  109. 109. Long GV, Fung C, Menzies AM, Pupo GM, Carlino MS, Hyman J, et al. Increased MAPK reactivation in early resistance to dabrafenib/trametinib combination therapy of BRAF-mutant metastatic melanoma. Nat Commun. 2014 Dec 2;5(1):5694
  110. 110. Marampon F, Ciccarelli C, Zani BM. Biological Rationale for Targeting MEK/ERK Pathways in Anti-Cancer Therapy and to Potentiate Tumour Responses to Radiation. Int J Mol Sci [Internet]. 2019 May 23 [cited 2021 Feb 22];20(10). Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6567863/
  111. 111. Qin J, Xin H, Nickoloff BJ. Specifically targeting ERK1 or ERK2 kills melanoma cells. J Transl Med. 2012 Jan 25;10:15
  112. 112. Little AS, Smith PD, Cook SJ. Mechanisms of acquired resistance to ERK1/2 pathway inhibitors. Oncogene. 2013 Mar;32(10):1207-1215

Written By

Dadnover Vargas-Ibarra, Mariana Velez-Vasquez and Maria Bermudez-Munoz

Submitted: 14 October 2020 Reviewed: 06 March 2021 Published: 12 April 2021