Next Article in Journal
The Role of Nutritional Status, Gastrointestinal Peptides, and Endocannabinoids in the Prognosis and Treatment of Children with Cancer
Previous Article in Journal
RNase 7 Inhibits Uropathogenic Escherichia coli-Induced Inflammation in Bladder Cells under a High-Glucose Environment by Regulating the JAK/STAT Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Discovery of Raf Family Is a Milestone in Deciphering the Ras-Mediated Intracellular Signaling Pathway

1
Queen Mary School, Nanchang University, Nanchang 330031, China
2
Provincial Key Laboratory of Tumor Pathogens and Molecular Pathology, Nanchang University, Nanchang 330031, China
3
NCU-QMUL Joint Research Institute of Precision Medical Science, Nanchang 330031, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(9), 5158; https://doi.org/10.3390/ijms23095158
Submission received: 17 March 2022 / Revised: 2 May 2022 / Accepted: 3 May 2022 / Published: 5 May 2022
(This article belongs to the Topic Kinases in Cancer and Other Diseases)

Abstract

:
The Ras-Raf-MEK-ERK signaling pathway, the first well-established MAPK pathway, plays essential roles in cell proliferation, survival, differentiation and development. It is activated in over 40% of human cancers owing to mutations of Ras, membrane receptor tyrosine kinases and other oncogenes. The Raf family consists of three isoforms, A-Raf, B-Raf and C-Raf. Since the first discovery of a truncated mutant of C-Raf as a transforming oncogene carried by a murine retrovirus, forty years of extensive studies have provided a wealth of information on the mechanisms underlying the activation, regulation and biological functions of the Raf family. However, the mechanisms by which activation of A-Raf and C-Raf is accomplished are still not completely understood. In contrast, B-Raf can be easily activated by binding of Ras-GTP, followed by cis-autophosphorylation of the activation loop, which accounts for the fact that this isoform is frequently mutated in many cancers, especially melanoma. The identification of oncogenic B-Raf mutations has led to accelerated drug development that targets Raf signaling in cancer. However, the effort has not proved as effective as anticipated, inasmuch as the mechanism of Raf activation involves multiple steps, factors and phosphorylation of different sites, as well as complex interactions between Raf isoforms. In this review, we will focus on the physiological complexity of the regulation of Raf kinases and their connection to the ERK phosphorylation cascade and then discuss the role of Raf in tumorigenesis and the clinical application of Raf inhibitors in the treatment of cancer.

1. Introduction

The Raf kinase family consists of three isoforms, C-Raf/Raf-1, B-Raf and A-Raf [1,2]. They are located directly downstream of Ras and upstream of MEK1/2 [3,4]. Since the discovery of v-Ras, v-Raf and ERK [5,6,7,8,9,10,11,12] and the connection of this regulatory hub with oncogenesis, tremendous efforts have been invested in the elucidation of the mechanisms underlying the activation of Raf kinases. The Raf/MEK/ERK pathway is the first and clearly defined mitogenic pathway whose signal is invoked by extracellular mitogenic ligands and serves as a framework for other MAPK pathways [13]. Among three isoforms of the Raf family, B-Raf is the only one that has so far been found mutated in many types of cancers [14]. As all of the Raf family members directly act downstream of Ras, C-Raf and A-Raf are also important factors contributing to oncogenesis, either mediating the effects of mutated Ras or participating in oncogenic B-Raf-mediated pathogenesis. Thus, the development of Raf inhibitors has been a focus in cancer therapy.
The Raf/MEK/ERK pathway plays important roles not only in physiological processes, including cell proliferation, differentiation and development, but also in oncogenesis and cancer progression [15]. The oncogenic Ras isoforms, KRas, HRas and NRas, whose mutations have been found in more than 30% of human cancers, as well as overexpression of growth factors and mutations of their receptors in human cancer, all lead to the activation of this pathway [11,12,16]. Furthermore, B-Raf mutations are present in approximately 8% of human cancers [17], including 50% of melanoma [14], 45% of papillary thyroid cancer [18], 10% of colon cancer [19], 10% of non-small cell lung cancer [20] and almost 100% of hairy cell leukemia cases [21]. Thus, the development of drugs targeting the Raf/MEK/ERK pathway is especially important for cancer therapy. To fulfill this objective, the elucidation of this pathway is a crucial step. Although the mechanisms of MEK and ERK activation are relatively straightforward, Raf activation is rather complex and still incompletely understood. Moreover, the discovery of homo- and hetero-dimerization of Raf isoforms adds to the complexity of deciphering the mechanism of their activation, which also makes the development of Raf inhibitors a challenging and daunting task. Therefore, this review will summarize current knowledge of the regulation of Raf isoforms and progress in the drug development of Raf inhibitors for cancer therapy.

2. Discovery of the Raf/MEK/ERK Pathway

In the early 1980s, v-Raf was initially identified as a transforming gene of the murine retrovirus 3611-murine sarcoma virus (MSV) [5,22]. In neonatal mice, it causes predominantly fibrosarcoma and erythroleukemias. Hence, the name of Raf originated from its capability to stimulate “Rapidly Accelerated Fibrosarcomas” [23]. Shortly after, the genome of the avian carcinoma virus MH2 was found to encode a closely related oncogene named v-mil [24]. Comparison of DNA sequences of these oncogenes coupled with biochemical studies has revealed that v-Raf and v-mil are retroviral oncogenes derived from cellular proto-oncogenes of mammalian and avian species. Both genes encode products that are classified into the serine/threonine kinase family, homologous to tyrosine specific SRC kinases in the kinase domain. v-Raf and v-mil are fused to the N-myristoylated (N-myr) viral Gag sequence but with the deletion of amino-terminal moieties, in contrast to their cellular counterparts [25,26].
The first cellular counterparts of v-Raf, C-Raf-1 and C-Raf-2, were cloned and sequenced in 1985. However, it soon became apparent that C-Raf-2 is a pseudogene. Thus, the C-Raf-1 gene product was designated as C-Raf. Studies have reported that C-Raf is located at human chromosomal band 3p25, consisting of nine exons that are similar to v-Raf and v-mil as well as two extra exons that are related to v-mil [27,28]. Later on, two more Raf family members, A-Raf and B-Raf, were identified in vertebrates [2,29,30]. Although C-Raf was the first to be discovered in mammals, the ortholog of B-Raf is D-Raf in Drosophila melanogaster and LIN-45 in Caenorhabditis elegans [31,32]. A-Raf is the smallest subtype with a molecular weight of 68 kDa, C-Raf having a weight of 73 kDa, while B-Raf ranges from 75 to 100 kDa, this being attributable to variable splicing [33,34].
The Human Protein Atlas expression database shows that C-Raf mRNA and A-Raf mRNA are predominantly present in skeletal muscle, bone marrow and the proximal digestive tract, while B-Raf mRNA is highly expressed in the retina, bone marrow and brain. Genetic and biochemical studies in C. elegans and D. melanogaster have demonstrated that Raf functions downstream of Ras and participates in cell proliferation, differentiation and development [31,32].
The mammalian mitogen-activated protein kinase (MAPK) was first identified in mammalian cells and then in yeast [35,36]. When adipocytes were briefly treated with insulin, a soluble serine/threonine kinase was rapidly activated, leading to increased phosphotyrosine content on microtubule-associated protein-2. Hence, it was initially named microtubule-associated protein-2 protein kinase (MAP-2 kinase). As it promotes cell cycle progression in response to insulin, growth factors and transforming proteins of oncogenic viruses, and phosphorylates a variety of protein substrates, it was renamed mitogen-activated protein kinase (MAPK) or extracellular signal-regulated kinase (ERK) [35,37,38]. Two isoforms, ERK1 and ERK2, are encoded by the MAPK3 and MAPK1 genes, respectively. Activation of ERK1/2 requires phosphorylation at threonine and tyrosine residues and inactivation through dephosphorylation [13,37,39,40]. Therefore, it is believed that a specific upstream kinase is responsible for phosphorylation and activation of ERK1/2 [41]. Subsequently, this kinase was found to be a dual-specificity kinase that phosphorylates threonine and tyrosine on ERK and is activated by nerve growth factor or epidermal growth factor, named MAP kinase kinase (MKK) or MAPK/ERK kinase (MEK) [42,43]. The kinase were cloned by independent groups [44,45]. Like ERK, MEK is also under negative regulation by dephosphorylation [13]. Further studies of the kinase cascade have revealed that Raf is an upstream kinase that controls MEK activity by phosphorylation. Thus, Raf kinase was positioned as the first serine/threonine upstream kinase of the canonical MEK-ERK pathway in 1992 [3] and a direct effector of Ras in 1993 [4,46,47].

3. The Mechanism of Raf Activation

3.1. Ras and 14-3-3

The Raf family possesses a similar structure consisting of three highly conserved regions, two (CR1 and CR2) in the N terminus [48] and the third (CR3) in the C terminus (Figure 1) [49]. CR1 contains a Ras-binding domain (RBD) which binds to Ras-GTP and a Cys-rich domain (CRD), the second Ras-binding site [50,51,52,53]. CR2 comprises a Ser/Thr-rich region, which bears important inhibitory phosphorylation sites, participating in the negative regulation of Ras binding and Raf activation. CR3 is the kinase domain [54]. As for C-Raf, truncation of the N-terminal region confers its ability to transform cells because of the elimination of inhibition imposed by the regulatory domains [55]. The overall homology of amino acid sequences between B-Raf and C-Raf kinase domains is 76%, and 74% between B-Raf and A-Raf kinase domains [29].
In a quiescent state, Raf maintains an inactive conformation through inter- and intra-molecular interactions. When growth factors bind to their cell-surface receptor tyrosine kinases (RTKs), growth-factor receptor-bound 2 (GRB2) and guanine nucleotide exchange factors such as Son of Sevenless (SOS) are recruited to the plasma membrane, which allows the exchange of GDP for GTP on Ras and initiates the process of Raf activation. In addition, G protein-coupled seven transmembrane receptors (GPCRs) can activate Ras [5,6,7,8,9,10,11,12] (Figure 2).
Ras is constitutively situated at the plasma membrane through prenylation of the carboxyterminal [56]. Upon loading with GTP, Ras interacts with Raf at two sites. First, it binds to the Ras binding domain (RBD) on Raf in a GTP-dependent manner [57,58]. Second, it binds to the cysteine-rich domain (CRD) independently of GTP [16,53,59]. This second binding stabilizes the interaction between Ras-GTP and RBD. In addition, CRD interacts with phosphatidyl-serine to locate Raf at the inner leaflet of the membrane [52,60,61].
An essential role is played by 14-3-3 in the regulation of Raf kinase activity [62,63,64,65]. There are two 14-3-3 binding sites, S259 and S621 (referring to C-Raf). While binding of 14-3-3 to pS259 exerts an inhibitory effect on Raf kinase activity [66], its binding to pS621 is essential for the kinase activity [62]. Therefore, it is postulated that the binding of a dimeric 14-3-3 to these two sites holds Raf in an inactive conformation. Ras-GTP binding destabilizes the association of 14-3-3 with pS259. As a result, one protomer of the dimeric 14-3-3 still sits on pS621, while the released one binds to pS621 of another molecule of Raf, resulting in the dimerization of Raf for further activation [67,68]. Almost at the same time, pS259 is dephosphorylated by phosphatases PP1 or PP2A to secure dimeric Raf at the plasma membrane [68,69,70].

3.2. The Role of Dimerization

Ras nanoclustering at the plasma membrane is driven by lipidated proteins at the inner plasma membrane, where the actin cytoskeleton is engaged in the generation of cholesterol-dependent clusters [71,72]. Thus, the formation of Ras nanoclusters promotes Raf dimerization/oligomerization. The first evidence for the importance of dimerization came from the observation that artificial dimerization of Raf strongly induces kinase activation [73,74]. Several studies have shown that in mammalian cells Raf family members form both homo- and hetero-dimeric complexes under physiological conditions [75,76,77,78]. However, the kinase activity of the B-Raf–C-Raf heterodimer is greater than that of the homodimer [75,76,77].

3.3. Regulation of Raf Kinase by Phosphorylation

The intrinsic kinase activity of C-Raf and A-Raf is tightly controlled until they are transported to the plasma membrane [79,80]. However, membrane recruitment is only the first step and is insufficient to stimulate their activation [81]. The membrane targeting brings Raf in close contact with activating kinases, including SRC family kinases and casein kinase 2 (CK2), which phosphorylate the activation loop and the N-terminal acidic region (NtA-region) between the N-terminal and C-terminal portions [82,83].
The existence of multiple phosphorylation sites reflects the fact that Raf proteins are subject to complex regulation. Phosphorylation has been documented to have both positive and negative impacts on Raf kinase activity; the impacts are listed in Table 1. Some phosphorylation sites are conserved across all three Raf family members, while others are isoform-specific, indicating common and distinct regulatory mechanisms [83]. As shown in Figure 1, S259 phosphorylated by PKB and possibly PKA plays a negative role after docking of 14-3-3, while binding of 14-3-3 to phosphorylated S621 is essential for Raf kinase activity [67,84]. However, Mischak et al. reported that phosphorylation of S621 exerted a negative impact on Raf kinase activity [85]. The data collected are based on in vitro studies. It is not clear whether they were performed in the presence or absence of 14-3-3, which might account for the discrepancy.

3.3.1. Positive Regulation

An SSYY motif (residues 338–341) in the NtA-region of C-Raf is conserved in A-Raf, which requires phosphorylation of both Ser338 and Tyr341 for their activation [86]. In the corresponding NtA-region of B-Raf (residues 446–449, SSDD), two tyrosine residues are replaced with aspartic acids and thus only S446 is phosphorylated, forming a salt bridge that stabilizes the Raf dimer [79,87,88]. These findings could explain why B-Raf has a higher basal activity than A-Raf and C-Raf.
S338 phosphorylation is usually used as a surrogate marker for C-Raf activation. p21-activated protein kinase (PAK) family kinases are reported to phosphorylate S338 in response to growth factor stimulation [88,89] and integrin activation as part of a PI3K–CDC42 and RAC signaling axis [89,90,91]. However, studies by us and others have indicated that PAK does not have a role in Ras-mediated induction of S338 phosphorylation [92], although the involvement of PAK in the phosphorylation and its association with C-Raf could be detected in the presence of nocodazole [92,93]. Similarly, several studies have reported that this site can be transphoshorylated in a heterodimer with B-Raf [75,76,77]. Another study has suggested that S338 is phosphorylated by MEK [75,76,77]. Our study suggested that S338 can be autophosphorylated in the homodimer form [78].
In addition to the NtA-region, phosphorylation of the activation loop is also critical to Raf activation [77,94,95]. Zhang et al. were the first to identify two phosphorylation sites in the activation loop, T599 and S602 on B-Raf, which are induced by Ras. Mutation of these two sites to alanine diminishes the kinase activity, while phospho-mimetic mutation (B-RafED) enhances the activity [95]. These two residues are conserved in other Raf isoforms, T491 and S494 in C-Raf, and T452 and T455 in A-Raf. Additionally, direct activation of C-Raf by PKC was observed via phosphorylation of S497/S499 [96,97]. The fact that the V600E mutation has been frequently found on B-Raf in some human cancers further reinforces the importance of phosphorylation of the activation loop.

3.3.2. Negative Regulation

Several studies have shown that S43 is phosphorylated by PKA, leading to impediment of Ras binding and C-Raf activation [98,99,100]. However, there are studies showing that PKA regulates other sites in the catalytic domain of C-Raf in addition to S43, possibly through phosphorylation of S621 [85,101,102]. Other phosphorylation sites include S289/S296/S301, which was reported by Balan et al. [103,104] and was found to play a positive role in mediating MEK/ERK feedback regulation. By contrast, S29/S43/S289/S296/S301/S642 were reported by Dougherty et al. [103] to be the phosphorylation sites that were responsive to MEK activation and which negatively regulated Raf kinase activation. Noticeably, some sites are the same in these two studies. The reasons underlying the discrepancies are not clear. Neither of these studies examined the role of individual phosphorylation sites but instead induced bulky mutations at all sites to alanine and then assessed their massive role in the regulation of kinase activity. Phosphorylation of B-Raf S151 reduces the dimerization of the kinases and also leads to direct disruption of B-Raf–C-Raf heterodimers [103].
Table 1. Impact of phosphorylation on C-Raf kinase activity.
Table 1. Impact of phosphorylation on C-Raf kinase activity.
SiteImpact on Raf Kinase ActivityKinaseReferences
S29NegativeKinases downstream of MEK1/2[103]
S43NegativePKA[98,99,100,105]
S259Negative,
14-3-3 binding
PKB, PKA[106,107,108,109,110]
S269PositiveKSR[111,112]
S289Negative, positiveKinases downstream of MEK1/2[103,104]
S296Negative, positiveKinases downstream of MEK1/2[103,104]
S301Negative, positiveKinases downstream of MEK1/2[103,104]
S338PositivePAK3, Raf, MEK[78,87,90,91]
Y341PositiveSrc[78,87,90,91]
S471Positive [77,94,95]
S497PositivePKC[96,97,113]
S499PositivePKC[96,97,113]
T491PositiveRaf or unclear[49]
S494PositiveRaf or unclear[49]
S621Negative or positive, 14-3-3 bindingRaf, PKA[67,84,85]
S642NegativeKinases downstream of MEK1/2[103]
With regard to the connection between phosphorylation and dimerization in Raf activation, several models have been proposed based on experimental data. For example, phosphorylation of Y341 facilitates that of S338; although phosphorylation of both sites does not require C-Raf dimerization in advance, the phosphorylation of Y341 promotes dimerization [114]. Recruitment of C-Raf to the plasma membrane depends on Ras binding but not on Raf dimerization. Shaw and his colleagues proposed a model in which allosteric Raf activation occurs in functionally asymmetric dimers [75,76,77]. In this model, B-Raf is first recruited to the plasma membrane and activated by Ras-GTP and then dimerized with C-Raf, leading to cis-autophosphorylation of C-Raf at the activation loop, which then phosphorylates and activates MEK. Finally, activated MEK induces phosphorylation of S338, resulting in full activation of C-Raf. In line with this, a study showed that MEK1 could activate C-Raf [115]. As for B-Raf, the NtA-region contains two aspartic acids (D448/D449) and S446 is constitutively phosphorylated [87]. Thus, this highly acidic region could promote homo- and hetero-dimerization. Of note, there are still questions that need to be addressed if the model holds. For example, tissue distribution of B-Raf and C-Raf is different, and deletion of their alleles generates distinct phenotypes. If their functions are always tied together, these aspects should be similar. Hence, it is conceivable that homo- and hetero-dimers exist in cells and execute different roles in qualitative and quantitative manners.
Raf kinases are subject to additional negative regulation such that physiological processes are well under control. If such feedback regulation is disabled, disordered biological consequences ensue, such as senescence and carcinogenesis [116,117,118]. Certainly, feedback inhibition is complex and variable, including with respect to direct and indirect control. For example, accumulation of ERK in the nucleus promotes expression of Raf kinase inhibitor protein (RKIP) which binds to Raf-1, MEK or ERK, interfering with key steps in activation of the pathway [119,120]. Another molecule that can interfere with the activation of downstream effectors is sprouty, which disrupts Ras–Raf interaction [121,122].
Activated Raf recruits and phosphorylates MEK1/2 at S218 and S222 in the activation loop. The phosphorylated MEK is released from the Raf–MEK complex and in turn phosphorylates ERK1/2 at conserved Threonine and Tyrosine residues in the activation loop, leading to activation of ERK1/2. The latter subsequently phosphorylates protein substrates in the cytoplasm and is also translocated to the nucleus to phosphorylate and regulate transcription factors [57,123,124]. This completes a canonical cascade of kinase at three levels, triggering cell-specific responses [125]. The three Raf isoforms have different abilities to activate MEK1 and MEK2. B-Raf is the strongest MEK kinase and A-Raf is the weakest MEK activator, which preferentially activates MEK1, while C-Raf has almost the same activity toward MEK1 and MEK2 [79,126]. In addition, Raf–MEK coupling is also promoted by PAK1 phosphorylation of MEK1 at S298 [127,128]. Since the linear ERK pathway was first delineated, many other molecules have been documented as being involved in the regulation of this pathway through crosstalk that entails positive and negative feedback mechanisms [129].

3.4. Scaffolds as Raf Regulators

Genetic screens in C. elegans and D. melanogaster have identified additional factors that contribute to Ras–ERK signaling. In addition to direct interaction between components of the Ras–ERK pathway, scaffold proteins play important roles in tethering them together, enabling efficient signal transmission. One of them is Kinase Suppressor of Ras (KSR), a kinase domain-containing protein [130,131,132]. Other scaffolding proteins include Connector Enhancer of KSR (CNK) [133], SUR-8 in C. elegans, known as SHOC2 in humans [134], β-arrestin [135], paxillin [136] and MAPK Organizer Protein 1 (MORG1) [137]. These proteins modulate the activation of Raf by Ras and physically bridge Raf to other components downstream of Ras, consequently facilitating pathway crosstalk.
Two KSR isoforms, KRS1 and KRS2, are found in mammals, which contain two distinctive regions, conserved area 1 (CA1) and conserved area 2 (CA2) but which lack RBD [132]. It is widely accepted that the primary function of KSR is to act as a scaffold to regulate the intensity and duration of the ERK pathway [138]. A study has suggested that KSR proteins act as allosteric inducers of Raf catalytic function [139]. This finding is in line with the observation that deletion of KSR renders mice resistant to tumor induction [140]. Other scaffolding activities of KSR have recently been reported, including crosstalk in Ras–ERK, calcium–calcineurin, and PKA signaling via phosphorylation-based regulation of the N- and C-terminal 14-3-3-binding sites on Raf [141,142].

4. Role of Raf in Biology

Genetic knockouts of different Raf isoforms in mice all lead to embryonic lethality or severe growth retardation and abnormal development [143,144,145]. Thus, A-Raf knockout mice were born alive but showed severe intestinal distension and neurological defects and died around postnatal day 20 [143]. B-Raf knockout mice succumbed in utero at embryonic day 12.5 due to massive bleeding in the body cavity, with severe vascular and neuronal abnormalities [144]. The phenotypes are attributed to increased apoptosis of endothelial cells and endothelial precursor cells in embryos and large blood vessels [146]. Ablation of C-Raf is fatal from embryonic day 10.5 to day 12.5, with dysplasia of the placenta, liver, hematopoietic organs, profound deafness and increased apoptosis of tissue cells [145,147,148,149]. It has been shown that C-Raf plays a critical role in the regulation of cell proliferation and suppression of apoptosis during embryogenesis [149]. A-Raf acts as a B-Raf effector and participates in Ras signaling when C-Raf is exhausted [150]. Moreover, A-Raf stabilizes B-Raf–C-Raf interaction to maintain signaling efficiency, especially in the presence of Raf inhibitors [150]. However, the cellular function of individual Raf is still less clear. Complex connections and interactions exist among them, which may depend on cell types or developmental stages.
Raf kinase family kinases play important roles in oncogenesis, inasmuch as they act as key effectors downstream of Ras, whose mutations account for oncogenesis in approximately 30% of human cancers, and as mediator for other oncogenes [11,150,151,152]. B-Raf has attracted great interest since the report that it was found to be mutated in 66% of malignant melanomas in 2002 [153]. Over 100 mutations in B-Raf have been identified in cancer patients. Most B-Raf mutations are concentrated in two regions: the glycine-rich P loop of the N lobe, and the activation segment and flanking regions in the kinase domain [154]. Among them, the most common mutation is a single amino acid substitution of valine 600 (V600, some reports designated V599) for glutamic acid, accounting for up to 90% of the mutations [124,153,155]. While most of the mutations significantly increase the kinase activity, some (e.g., B-Raf G595R) exhibit impaired activity and cannot phosphorylate MEK directly [156]. However, the B-Raf mutants with decreased kinase activity could hyper-stimulate the ERK pathway [154,157]. Therefore, the oncogenic mechanism of B-Raf is fundamentally different from that of the constitutively activated v-Raf found in murine retrovirus [158].
Although mutations of Raf-1 are much rarer in cancer than B-Raf, several studies have reported germline mutations of C-Raf in human diseases. For example, two mutations, S427G and I448V, are found in the kinase domain of C-Raf in patients with therapy-related acute myeloid leukemia [159]. The mutation of S427G causes increased activity of the Raf/MEK/ERK pathway, while I448V mutation does not affect the kinase activity. This study suggests that these germline mutations of C-Raf are predisposing factors for human neoplasia. In addition, mutation of C-Raf has been documented in “RASopathies”—a diverse collection of disorders caused by germline mutations in genes that code for the components or regulators of the RAS-RAF-MEK-ERK pathway. The disorders are characterized by postnatal short stature and neurocognitive delay, including neurofibromatosis type 1, Noonan syndrome, Noonan syndrome with multiple lentigines, cardio-facio-cutaneous syndrome and Legius syndrome [160,161,162].
In terms of dimerization state, there are three major types of B-Raf mutations described in human cancers (Figure 3). The most common one is the Class I with V600E mutation, which renders the kinase active as a monomer. The mutation mimics phosphorylation of the activation loop, leading to disruption of inactive conformation [153]. Class II mutations, including K601E, L597Q, and G469A, cause spontaneous dimerization, resulting in the activation of the kinase. These mutations destabilize auto-inhibition by disrupting the inhibitory interaction of the activation loop with the Gly-rich loop and disable feedback suppression of Raf dimers [163]. Class III mutations impair kinase activity toward MEK and adopt a tumor-specific mechanism by which the mutants transactivate endogenous C-Raf through phosphorylation of the activation loop by forming B-Raf–C-Raf heterodimers [164]. Unlike Class I and Class II mutants, Class III mutants bind to Ras more tightly than wild-type B-Raf. Intriguingly, mutations equivalent to B-Raf V600E in C-Raf and A-Raf fail to produce oncogenic effects unless a negative charge is introduced into the NtA-region. These results support the notion that the regulation of C-Raf and A-Raf kinase domains is tighter because mutations at the two sites are required to confer transforming ability [163,165].

5. Development of Raf Inhibitors

Due to the prevalence of B-Raf mutations in melanomas, the development of Raf inhibitors has become a research hotspot (Table 2). To understand the mechanism by which Raf kinase inhibitors function, Lavoie and Therrien [1] proposed a model to illustrate the mechanism underlying the actions of Raf inhibitors (Figure 4). Wild-type Raf maintains an inactive conformation featuring an αC-helix and a DFG OUT position. Dimerization of Raf results in a closed conformation between the N-terminal lobe (N-lobe) and C-terminal lobe (C-lobe) with an αC-helix and DFG transitioned into the IN position to stabilize the side-to-side interface [166,167]. Exceptionally, V600E mutation leads to B-Raf activation even in the form of a monomer. Raf inhibitors can stabilize the αC-helix at a position between IN and OUT conformation, forming defective dimers due to variances in the position of each protomer αC-helix. Hence, Raf inhibitors are divided into two categories according to the conformational types of Raf they act on. One class consists of the ‘αC-IN’ inhibitors; the other consists of the ‘αC-OUT’ inhibitors, such as vemurafenib and dabrafenib [168].
Both the first and second generations of Raf inhibitors bind to Raf in an ATP-competitive manner but recognize different conformations. The first generation of Raf inhibitors consists of small ATP-competitive ‘αC-IN’ inhibitors (refer to Figure 4), such as Sorafenib [169]. These can faintly inhibit the monomer activity of B-Raf V600E. The major limitation is that the first generation of Raf inhibitors cannot inhibit wild-type Raf but instead promote dimerization of Raf, resulting in the transactivation of wild-type Raf by V600E mutants of B-Raf, followed by activation of the ERK1/2 pathway. This phenomenon is known as the ‘B-Raf inhibitor paradox’ [15,170,171].
The second generation of Raf inhibitors, including vemurafenib and dabrafenib, consists of ‘αC-OUT’ Raf inhibitors approved by the FDA for clinical use [172,173]. These two drugs are relatively unable to inhibit dimeric Raf in non-Raf mutated cells but are effective in inhibiting B-Raf V600E and can lead to subsequent development of dimer-driven resistance. They are predicted not to be effective in tumors with non-V600 mutant B-Raf [172,174]. In most cases, the drugs are limited by the development of drug resistance and tumor recurrence, although they extend patient life spans to some extent. In addition to acquired resistance, increased Raf dimerization may lead to an adaptive response to Raf inhibitor therapy, as increased Ras activation may promote Raf dimerization [168,175]. However, it has been found that increasing the concentration of inhibitors to occupy two dimer partners could overcome this paradoxical activation mechanism [176].
Development of the third generation of Raf inhibitors is underway to solve problems associated with the paradoxical activation arising from the previous two generations of Raf inhibitors. According to structural and biochemical studies, they are divided into two categories, pan-Raf inhibitors and paradox breakers (Figure 5). Pan-Raf inhibitors include AZ628, belvarafenib, CCT196969, CCT241161, LY3009120 and TAK-580 (MLN2480); these target ‘DFG-OUT’ and ‘αC-IN’ conformations of Raf [177,178,179,180,181,182,183]. Binding of active Raf dimers and monomers with similar affinity leads to inhibition of ERK signaling in cells containing active Raf monomers or dimers. Although they could potentiate the dimerization of Raf by stabilizing ‘αC-IN’ conformation, transactivation is prevented due to their similar binding affinity for and inhibition of wild-type B-Raf or C-Raf. Of note, although the pan-Raf inhibitors are effective in vitro, due to their lack of selectivity for B-Raf mutations, they also inhibit wild-type Raf dimers in normal cells [183,184,185]. Hence, their application in vivo is more limited, which significantly reduces their therapeutic index.
Paradox breakers are ATP-competitive inhibitors and include PLX7904 and its analogue PLX8394 [177,186,187]. This class of inhibitor binds closely to Leu505, which disrupts the Raf dimer interface and allosterically blocks the kinase activity. They can specifically perturb B-Raf dimers, including B-Raf homodimers and B-Raf–C-Raf heterodimers that commonly exist in oncogenic Ras-containing cells, but not C-Raf homodimers or A-Raf homodimers (Figure 5). This novel dimer selective inhibitor neither activates nor inhibits wild-type Raf and will have a wider therapeutic window than pan-Raf inhibitors [177]. However, these drugs are ineffective against resistance resulting from Ras activation or by any other means of activation of C-Raf homodimers [188,189,190].
In general, the Raf inhibitors that have been approved so far only block B-Raf V600 mutant monomers, rendering them ineffective against malignancies where Raf signals as a dimer. The pan-Raf inhibitors have equal affinity for both protomers in Raf dimers and have less selectivity within Raf isoforms. Thus, they may not only suppress mutant dimers in tumors but also impair MAPK signaling in normal cells. Phase I research on the paradox breakers is presently underway. They particularly disrupt B-Raf dimers, which might help to overcome the limitations of pan-Raf inhibitors and improve therapeutic outcomes.
In melanomas containing B-Raf V600E mutations, Raf inhibitors are used in combination with MEK inhibitors to provide more effective and durable inhibition of ERK signaling [191,192]. As a result, the combination of drugs delays the development of drug resistance and reduces the toxicity associated with B-Raf inhibitors seen in monotherapy. The improved efficacy of the combination is associated with the ability of MEK inhibitors to prevent the reactivation of residual ERK. Three MEK inhibitors have been approved by the FDA in combination with B-Raf inhibitors—trametinib, cobimetinib and binimetinib. For example, combined use of dabrafenib and trametinib increases the response rate from 54% to 76% and the median duration of response from 5.8 months to 10.5 months [193,194,195,196]. As pan-Raf inhibitors, MEK inhibitors and ERK inhibitors block the MAPK pathway in normal cells; the therapeutic window for these agents is narrower than that for B-Raf kinase inhibitors, such as vemurafenib, dabrafenib and encorafenib, which limits their clinical use. Several next generation ‘αC-IN’ Raf inhibitors designed to inhibit dimeric Raf and thus be more effective in a wide range of tumors containing non-V600E B-Raf or activating mutants of Ras are currently in clinical trials [15].

6. Conclusions

Since the discovery of v-Raf, great efforts have been made to elucidate the signaling transduction pathway involving Raf and its role in physiological and pathophysiological functions. Although the framework of the Ras/Raf/MEK/ERK pathway has been delineated, the mechanisms of Raf activation, especially C-Raf and A-Raf, which involve other factors, kinases and inter-regulation among Raf isoforms, are not fully understood. This being so, it is difficult to develop Raf kinase inhibitors as cancer therapeutic agents. Although several inhibitors of Raf are used in clinic or are in clinical trials, resistance is quickly developed, which limits their use. In the future, more effective drugs will rely on the context of B-Raf mutations as a result of the unravelling of the complexity of Raf isoform interaction and inter-regulation. The goal of developing Raf inhibitors is to eliminate cancerous cells but not normal cells.

Author Contributions

Conceptualization, J.Z. and Z.L.; writing—original draft preparation, J.Z.; writing—review and editing, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

The Innovation and Entrepreneurship grant from the Jiangxi Province Bureau of Foreign Experts.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful to Yun He for her suggestions on preparation of Figures.

Conflicts of Interest

We hereby declare that all authors have no conflict of interest in this work, including affiliations, financial relationships, personal relationships or funding sources that could be perceived as influencing an author’s objectivity regarding the manuscript content (2014).

References

  1. Lavoie, H.; Therrien, M. Regulation of RAF protein kinases in ERK signalling. Nat. Rev. Mol. Cell. Biol. 2015, 16, 281–298. [Google Scholar] [CrossRef]
  2. Matallanas, D.; Birtwistle, M.; Romano, D.; Zebisch, A.; Rauch, J.; von Kriegsheim, A.; Kolch, W. Raf family kinases: Old dogs have learned new tricks. Genes Cancer 2011, 2, 232–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kyriakis, J.M.; App, H.; Zhang, X.F.; Banerjee, P.; Brautigan, D.L.; Rapp, U.R.; Avruch, J. Raf-1 activates MAP kinase-kinase. Nature 1992, 358, 417–421. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, X.F.; Settleman, J.; Kyriakis, J.M.; Takeuchi-Suzuki, E.; Elledge, S.J.; Marshall, M.S.; Bruder, J.T.; Rapp, U.R.; Avruch, J. Normal and oncogenic p21ras proteins bind to the amino-terminal regulatory domain of c-Raf-1. Nature 1993, 364, 308–313. [Google Scholar] [CrossRef] [PubMed]
  5. Rapp, U.R.; Goldsborough, M.D.; Mark, G.E.; Bonner, T.I.; Groffen, J.; Reynolds, F.H., Jr.; Stephenson, J.R. Structure and biological activity of v-raf, a unique oncogene transduced by a retrovirus. Proc. Natl. Acad. Sci. USA 1983, 80, 4218–4222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Brown, K.; Quintanilla, M.; Ramsden, M.; Kerr, I.B.; Young, S.; Balmain, A. v-ras genes from Harvey and BALB murine sarcoma viruses can act as initiators of two-stage mouse skin carcinogenesis. Cell 1986, 46, 447–456. [Google Scholar] [CrossRef]
  7. Shimizu, K.; Goldfarb, M.; Suard, Y.; Perucho, M.; Li, Y.; Kamata, T.; Feramisco, J.; Stavnezer, E.; Fogh, J.; Wigler, M.H. Three human transforming genes are related to the viral ras oncogenes. Proc. Natl. Acad. Sci. USA 1983, 80, 2112–2116. [Google Scholar] [CrossRef] [Green Version]
  8. Ladeda, V.; Frankel, P.; Feig, L.A.; Foster, D.A.; Bal de Kier Joffe, E.; Aguirre-Ghiso, J.A. RalA mediates v-Src, v-Ras, and v-Raf regulation of CD44 and fibronectin expression in NIH3T3 fibroblasts. Biochem. Biophys. Res. Commun. 2001, 283, 854–861. [Google Scholar] [CrossRef]
  9. Pisapia, P.; Pepe, F.; Iaccarino, A.; Sgariglia, R.; Nacchio, M.; Russo, G.; Gragnano, G.; Malapelle, U.; Troncone, G. BRAF: A Two-Faced Janus. Cells 2020, 9, 2549. [Google Scholar] [CrossRef]
  10. Leevers, S.J.; Marshall, C.J. Activation of extracellular signal-regulated kinase, ERK2, by p21ras oncoprotein. EMBO J. 1992, 11, 569–574. [Google Scholar] [CrossRef]
  11. Cox, A.D.; Der, C.J. Ras history: The saga continues. Small GTPases 2010, 1, 2–27. [Google Scholar] [CrossRef] [Green Version]
  12. Malumbres, M.; Barbacid, M. RAS oncogenes: The first 30 years. Nat. Rev. Cancer 2003, 3, 459–465. [Google Scholar] [CrossRef]
  13. Ahn, N.G. The MAP kinase cascade. Discovery of a new signal transduction pathway. Mol. Cell. Biochem. 1993, 127–128, 201–209. [Google Scholar] [CrossRef]
  14. Akbani, R.; Akdemir, K.C.; Aksoy, B.A.; Albert, M.; Ally, A.; Amin, S.B.; Kwong, L.N. Genomic Classification of Cutaneous Melanoma. Cell 2015, 161, 1681–1696. [Google Scholar] [CrossRef] [Green Version]
  15. Karoulia, Z.; Gavathiotis, E.; Poulikakos, P.I. New perspectives for targeting RAF kinase in human cancer. Nat. Rev. Cancer 2017, 17, 676–691. [Google Scholar] [CrossRef]
  16. Hancock, J.F. Ras proteins: Different signals from different locations. Nat. Rev. Mol. Cell. Biol. 2003, 4, 373–384. [Google Scholar] [CrossRef]
  17. Garnett, M.J.; Marais, R. Guilty as charged: B-RAF is a human oncogene. Cancer Cell 2004, 6, 313–319. [Google Scholar] [CrossRef] [Green Version]
  18. Kimura, E.T.; Nikiforova, M.N.; Zhu, Z.; Knauf, J.A.; Nikiforov, Y.E.; Fagin, J.A. High prevalence of BRAF mutations in thyroid cancer: Genetic evidence for constitutive activation of the RET/PTC-RAS-BRAF signaling pathway in papillary thyroid carcinoma. Cancer Res. 2003, 63, 1454–1457. [Google Scholar] [CrossRef] [Green Version]
  19. Rajagopalan, H.; Bardelli, A.; Lengauer, C.; Kinzler, K.W.; Vogelstein, B.; Velculescu, V.E. Tumorigenesis: RAF/RAS oncogenes and mismatch-repair status. Nature 2002, 418, 934. [Google Scholar] [CrossRef]
  20. Never-smoker, N.E.S. Comprehensive molecular profiling of lung adenocarcinoma. Nature 2014, 511, 543–550. [Google Scholar] [CrossRef]
  21. Tiacci, E.; Trifonov, V.; Schiavoni, G.; Holmes, A.; Kern, W.; Martelli, M.P.; Pucciarini, A.; Bigerna, B.; Pacini, R.; Wells, V.A.; et al. BRAF mutations in hairy-cell leukemia. N. Engl. J. Med. 2011, 364, 2305–2315. [Google Scholar] [CrossRef] [Green Version]
  22. Rapp, U.R.; Reynolds, F.H., Jr.; Stephenson, J.R. New mammalian transforming retrovirus: Demonstration of a polyprotein gene product. J. Virol. 1983, 45, 914–924. [Google Scholar] [CrossRef] [Green Version]
  23. Rapp, U.R.; Cleveland, J.L.; Fredrickson, T.N.; Holmes, K.L.; Morse, H.C., 3rd; Jansen, H.W.; Patschinsky, T.; Bister, K. Rapid induction of hemopoietic neoplasms in newborn mice by a raf(mil)/myc recombinant murine retrovirus. J. Virol. 1985, 55, 23–33. [Google Scholar] [CrossRef] [Green Version]
  24. Jansen, H.W.; Lurz, R.; Bister, K.; Bonner, T.I.; Mark, G.E.; Rapp, U.R. Homologous cell-derived oncogenes in avian carcinoma virus MH2 and murine sarcoma virus 3611. Nature 1984, 307, 281–284. [Google Scholar] [CrossRef]
  25. Rapp, U.R.; Todaro, C. Generation of new mouse sarcoma viruses in cell culture. Science 1978, 201, 821–824. [Google Scholar] [CrossRef] [PubMed]
  26. Moelling, K.; Heimann, B.; Beimling, P.; Rapp, U.R.; Sander, T. Serine- and threonine-specific protein kinase activities of purified gag-mil and gag-raf proteins. Nature 1984, 312, 558–561. [Google Scholar] [CrossRef] [PubMed]
  27. Bonner, T.I.; Kerby, S.B.; Sutrave, P.; Gunnell, M.A.; Mark, G.; Rapp, U.R. Structure and biological activity of human homologs of the raf/mil oncogene. Mol. Cell. Biol. 1985, 5, 1400–1407. [Google Scholar] [CrossRef] [PubMed]
  28. Bonner, T.; O’Brien, S.J.; Nash, W.G.; Rapp, U.R.; Morton, C.C.; Leder, P. The human homologs of the raf (mil) oncogene are located on human chromosomes 3 and 4. Science 1984, 223, 71–74. [Google Scholar] [CrossRef]
  29. Ikawa, S.; Fukui, M.; Ueyama, Y.; Tamaoki, N.; Yamamoto, T.; Toyoshima, K. B-raf, a new member of the raf family, is activated by DNA rearrangement. Mol. Cell. Biol. 1988, 8, 2651–2654. [Google Scholar] [CrossRef]
  30. Huleihel, M.; Goldsborough, M.; Cleveland, J.; Gunnell, M.; Bonner, T.; Rapp, U.R. Characterization of murine A-raf, a new oncogene related to the v-raf oncogene. Mol. Cell. Biol. 1986, 6, 2655–2662. [Google Scholar] [CrossRef]
  31. Han, M.; Golden, A.; Han, Y.; Sternberg, P.W.C. C. elegans lin-45 raf gene participates in let-60 ras-stimulated vulval differentiation. Nature 1993, 363, 133–140. [Google Scholar] [CrossRef]
  32. Ambrosio, L.; Mahowald, A.P.; Perrimon, N. Requirement of the Drosophila raf homologue for torso function. Nature 1989, 342, 288–291. [Google Scholar] [CrossRef]
  33. Barnier, J.V.; Papin, C.; Eychène, A.; Lecoq, O.; Calothy, G. The mouse B-raf gene encodes multiple protein isoforms with tissue-specific expression. J. Biol. Chem. 1995, 270, 23381–23389. [Google Scholar] [CrossRef] [Green Version]
  34. Storm, S.M.; Cleveland, J.L.; Rapp, U.R. Expression of raf family proto-oncogenes in normal mouse tissues. Oncogene 1990, 5, 345–351. [Google Scholar] [CrossRef]
  35. Sturgill, T.W.; Ray, L.B.; Erikson, E.; Maller, J.L. Insulin-stimulated MAP-2 kinase phosphorylates and activates ribosomal protein S6 kinase II. Nature 1988, 334, 715–718. [Google Scholar] [CrossRef]
  36. Courchesne, W.E.; Kunisawa, R.; Thorner, J. A putative protein kinase overcomes pheromone-induced arrest of cell cycling in S. cerevisiae. Cell 1989, 58, 1107–1119. [Google Scholar] [CrossRef]
  37. Ray, L.B.; Sturgill, T.W. Insulin-stimulated microtubule-associated protein kinase is phosphorylated on tyrosine and threonine in vivo. Proc. Natl. Acad. Sci. USA 1988, 85, 3753–3757. [Google Scholar] [CrossRef] [Green Version]
  38. Lindberg, R.A.; Quinn, A.M.; Hunter, T. Dual-specificity protein kinases: Will any hydroxyl do? Trends Biochem. Sci. 1992, 17, 114–119. [Google Scholar] [CrossRef]
  39. Anderson, N.G.; Maller, J.L.; Tonks, N.K.; Sturgill, T.W. Requirement for integration of signals from two distinct phosphorylation pathways for activation of MAP kinase. Nature 1990, 343, 651–653. [Google Scholar] [CrossRef]
  40. Ullah, R.; Yin, Q.; Snell, A.H.; Wan, L. RAF-MEK-ERK pathway in cancer evolution and treatment. Semin. Cancer Biol. 2021. [Google Scholar] [CrossRef]
  41. Dent, P.; Haser, W.; Haystead, T.A.; Vincent, L.A.; Roberts, T.M.; Sturgill, T.W. Activation of mitogen-activated protein kinase kinase by v-Raf in NIH 3T3 cells and in vitro. Science 1992, 257, 1404–1407. [Google Scholar] [CrossRef] [PubMed]
  42. Gómez, N.; Cohen, P. Dissection of the protein kinase cascade by which nerve growth factor activates MAP kinases. Nature 1991, 353, 170–173. [Google Scholar] [CrossRef] [PubMed]
  43. Ahn, N.G.; Seger, R.; Bratlien, R.L.; Diltz, C.D.; Tonks, N.K.; Krebs, E.G. Multiple components in an epidermal growth factor-stimulated protein kinase cascade. In vitro activation of a myelin basic protein/microtubule-associated protein 2 kinase. J. Biol. Chem. 1991, 266, 4220–4227. [Google Scholar] [CrossRef]
  44. Crews, C.M.; Alessandrini, A.; Erikson, R.L. The primary structure of MEK, a protein kinase that phosphorylates the ERK gene product. Science 1992, 258, 478–480. [Google Scholar] [CrossRef]
  45. Zheng, C.F.; Guan, K.L. Cloning and characterization of two distinct human extracellular signal-regulated kinase activator kinases, MEK1 and MEK2. J. Biol. Chem. 1993, 268, 11435–11439. [Google Scholar] [CrossRef]
  46. Vojtek, A.B.; Hollenberg, S.M.; Cooper, J.A. Mammalian Ras interacts directly with the serine/threonine kinase Raf. Cell 1993, 74, 205–214. [Google Scholar] [CrossRef]
  47. Warne, P.H.; Viciana, P.R.; Downward, J. Direct interaction of Ras and the amino-terminal region of Raf-1 in vitro. Nature 1993, 364, 352–355. [Google Scholar] [CrossRef]
  48. Heidecker, G.; Huleihel, M.; Cleveland, J.L.; Kolch, W.; Beck, T.W.; Lloyd, P.; Pawson, T.; Rapp, U.R. Mutational activation of c-raf-1 and definition of the minimal transforming sequence. Mol. Cell. Biol. 1990, 10, 2503–2512. [Google Scholar] [CrossRef]
  49. Chong, H.; Lee, J.; Guan, K.L. Positive and negative regulation of Raf kinase activity and function by phosphorylation. EMBO J. 2001, 20, 3716–3727. [Google Scholar] [CrossRef] [Green Version]
  50. Chuang, E.; Barnard, D.; Hettich, L.; Zhang, X.F.; Avruch, J.; Marshall, M.S. Critical binding and regulatory interactions between Ras and Raf occur through a small, stable N-terminal domain of Raf and specific Ras effector residues. Mol. Cell. Biol. 1994, 14, 5318–5325. [Google Scholar] [CrossRef]
  51. Beck, T.W.; Huleihel, M.; Gunnell, M.; Bonner, T.I.; Rapp, U.R. The complete coding sequence of the human A-raf-1 oncogene and transforming activity of a human A-raf carrying retrovirus. Nucleic Acids Res. 1987, 15, 595–609. [Google Scholar] [CrossRef] [Green Version]
  52. Tran, N.H.; Wu, X.; Frost, J.A. B-Raf and Raf-1 are regulated by distinct autoregulatory mechanisms. J. Biol. Chem. 2005, 280, 16244–16253. [Google Scholar] [CrossRef] [Green Version]
  53. Luo, Z.; Diaz, B.; Marshall, M.S.; Avruch, J. An intact Raf zinc finger is required for optimal binding to processed Ras and for ras-dependent Raf activation in situ. Mol. Cell. Biol. 1997, 17, 46–53. [Google Scholar] [CrossRef] [Green Version]
  54. Dhillon, A.S.; Meikle, S.; Yazici, Z.; Eulitz, M.; Kolch, W. Regulation of Raf-1 activation and signalling by dephosphorylation. EMBO J. 2002, 21, 64–71. [Google Scholar] [CrossRef]
  55. Ishikawa, F.; Takaku, F.; Nagao, M.; Sugimura, T. The complete primary structure of the rat A-raf cDNA coding region: Conservation of the putative regulatory regions present in rat c-raf. Oncogene Res. 1987, 1, 243–253. [Google Scholar] [CrossRef] [Green Version]
  56. Ahearn, I.M.; Haigis, K.; Bar-Sagi, D.; Philips, M.R. Regulating the regulator: Post-translational modification of RAS. Nat. Rev. Mol. Cell. Biol. 2011, 13, 39–51. [Google Scholar] [CrossRef] [Green Version]
  57. Yoon, S.; Seger, R. The extracellular signal-regulated kinase: Multiple substrates regulate diverse cellular functions. Growth Factors 2006, 24, 21–44. [Google Scholar] [CrossRef]
  58. Zhao, C.; Du, G.; Skowronek, K.; Frohman, M.A.; Bar-Sagi, D. Phospholipase D2-generated phosphatidic acid couples EGFR stimulation to Ras activation by Sos. Nat. Cell Biol. 2007, 9, 706–712. [Google Scholar] [CrossRef]
  59. Tran, T.H.; Chan, A.H.; Young, L.C.; Bindu, L.; Neale, C.; Messing, S.; Dharmaiah, S.; Taylor, T.; Denson, J.P.; Esposito, D.; et al. KRAS interaction with RAF1 RAS-binding domain and cysteine-rich domain provides insights into RAS-mediated RAF activation. Nat. Commun. 2021, 12, 1176. [Google Scholar] [CrossRef]
  60. Roy, S.; Lane, A.; Yan, J.; McPherson, R.; Hancock, J.F. Activity of plasma membrane-recruited Raf-1 is regulated by Ras via the Raf zinc finger. J. Biol. Chem. 1997, 272, 20139–20145. [Google Scholar] [CrossRef] [Green Version]
  61. Ghosh, S.; Strum, J.C.; Sciorra, V.A.; Daniel, L.; Bell, R.M. Raf-1 kinase possesses distinct binding domains for phosphatidylserine and phosphatidic acid. Phosphatidic acid regulates the translocation of Raf-1 in 12-O-tetradecanoylphorbol-13-acetate-stimulated Madin-Darby canine kidney cells. J. Biol. Chem. 1996, 271, 8472–8480. [Google Scholar] [CrossRef] [Green Version]
  62. Luo, Z.J.; Zhang, X.F.; Rapp, U.; Avruch, J. Identification of the 14.3.3 zeta domains important for self-association and Raf binding. J. Biol. Chem. 1995, 270, 23681–23687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Li, S.; Janosch, P.; Tanji, M.; Rosenfeld, G.C.; Waymire, J.C.; Mischak, H.; Kolch, W.; Sedivy, J.M. Regulation of Raf-1 kinase activity by the 14-3-3 family of proteins. EMBO J. 1995, 14, 685–696. [Google Scholar] [CrossRef]
  64. Fantl, W.J.; Muslin, A.J.; Kikuchi, A.; Martin, J.A.; MacNicol, A.M.; Gross, R.W.; Williams, L.T. Activation of Raf-1 by 14-3-3 proteins. Nature 1994, 371, 612–614. [Google Scholar] [CrossRef] [PubMed]
  65. Freed, E.; Symons, M.; Macdonald, S.G.; McCormick, F.; Ruggieri, R. Binding of 14-3-3 proteins to the protein kinase Raf and effects on its activation. Science 1994, 265, 1713–1716. [Google Scholar] [CrossRef]
  66. Jia, S.; Flores-Saaib, R.D.; Courey, A.J. The Dorsal Rel homology domain plays an active role in transcriptional regulation. Mol. Cell. Biol. 2002, 22, 5089–5099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Tzivion, G.; Luo, Z.; Avruch, J. A dimeric 14-3-3 protein is an essential cofactor for Raf kinase activity. Nature 1998, 394, 88–92. [Google Scholar] [CrossRef] [PubMed]
  68. Rommel, C.; Radziwill, G.; Lovrić, J.; Noeldeke, J.; Heinicke, T.; Jones, D.; Aitken, A.; Moelling, K. Activated Ras displaces 14-3-3 protein from the amino terminus of c-Raf-1. Oncogene 1996, 12, 609–619. [Google Scholar] [PubMed]
  69. Raabe, T.; Rapp, U.R. Ras signaling: PP2A puts Ksr and Raf in the right place. Curr. Biol. 2003, 13, R635–R637. [Google Scholar] [CrossRef] [Green Version]
  70. Jaumot, M.; Hancock, J.F. Protein phosphatases 1 and 2A promote Raf-1 activation by regulating 14-3-3 interactions. Oncogene 2001, 20, 3949–3958. [Google Scholar] [CrossRef]
  71. Plowman, S.J.; Muncke, C.; Parton, R.G.; Hancock, J.F. H-ras, K-ras, and inner plasma membrane raft proteins operate in nanoclusters with differential dependence on the actin cytoskeleton. Proc. Natl. Acad. Sci. USA 2005, 102, 15500–15505. [Google Scholar] [CrossRef] [Green Version]
  72. Tian, T.; Harding, A.; Inder, K.; Plowman, S.; Parton, R.G.; Hancock, J.F. Plasma membrane nanoswitches generate high-fidelity Ras signal transduction. Nat. Cell Biol. 2007, 9, 905–914. [Google Scholar] [CrossRef]
  73. Farrar, M.A.; Alberol-Ila, J.; Perlmutter, R.M. Activation of the Raf-1 kinase cascade by coumermycin-induced dimerization. Nature 1996, 383, 178–181. [Google Scholar] [CrossRef]
  74. Luo, Z.; Tzivion, G.; Belshaw, P.J.; Vavvas, D.; Marshall, M.; Avruch, J. Oligomerization activates c-Raf-1 through a Ras-dependent mechanism. Nature 1996, 383, 181–185. [Google Scholar] [CrossRef]
  75. Weber, C.K.; Slupsky, J.R.; Kalmes, H.A.; Rapp, U.R. Active Ras induces heterodimerization of cRaf and BRaf. Cancer Res. 2001, 61, 3595–3598. [Google Scholar]
  76. Rushworth, L.K.; Hindley, A.D.; O’Neill, E.; Kolch, W. Regulation and role of Raf-1/B-Raf heterodimerization. Mol. Cell. Biol. 2006, 26, 2262–2272. [Google Scholar] [CrossRef] [Green Version]
  77. Hu, J.; Stites, E.C.; Yu, H.; Germino, E.A.; Meharena, H.S.; Stork, P.J.S.; Kornev, A.P.; Taylor, S.S.; Shaw, A.S. Allosteric activation of functionally asymmetric RAF kinase dimers. Cell 2013, 154, 1036–1046. [Google Scholar] [CrossRef] [Green Version]
  78. Zang, M.; Gong, J.; Luo, L.; Zhou, J.; Xiang, X.; Huang, W.; Huang, Q.; Luo, X.; Olbrot, M.; Peng, Y.; et al. Characterization of Ser338 phosphorylation for Raf-1 activation. J. Biol. Chem. 2008, 283, 31429–31437. [Google Scholar] [CrossRef] [Green Version]
  79. Marais, R.; Light, Y.; Paterson, H.F.; Mason, C.S.; Marshall, C.J. Differential regulation of Raf-1, A-Raf, and B-Raf by oncogenic ras and tyrosine kinases. J. Biol. Chem. 1997, 272, 4378–4383. [Google Scholar] [CrossRef] [Green Version]
  80. Mineo, C.; Anderson, R.G.; White, M.A. Physical association with ras enhances activation of membrane-bound raf (RafCAAX). J. Biol. Chem. 1997, 272, 10345–10348. [Google Scholar] [CrossRef]
  81. Weber, C.K.; Slupsky, J.R.; Herrmann, C.; Schuler, M.; Rapp, U.R.; Block, C. Mitogenic signaling of Ras is regulated by differential interaction with Raf isozymes. Oncogene 2000, 19, 169–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Chow, Y.H.; Pumiglia, K.; Jun, T.H.; Dent, P.; Sturgill, T.W.; Jove, R. Functional mapping of the N-terminal regulatory domain in the human Raf-1 protein kinase. J. Biol. Chem. 1995, 270, 14100–14106. [Google Scholar] [CrossRef] [Green Version]
  83. Marais, R.; Light, Y.; Paterson, H.F.; Marshall, C.J. Ras recruits Raf-1 to the plasma membrane for activation by tyrosine phosphorylation. EMBO J. 1995, 14, 3136–3145. [Google Scholar] [CrossRef] [PubMed]
  84. Noble, C.; Mercer, K.; Hussain, J.; Carragher, L.; Giblett, S.; Hayward, R.; Patterson, C.; Marais, R.; Pritchard, C.A. CRAF autophosphorylation of serine 621 is required to prevent its proteasome-mediated degradation. Mol. Cell 2008, 31, 862–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Mischak, H.; Seitz, T.; Janosch, P.; Eulitz, M.; Steen, H.; Schellerer, M.; Philipp, A.; Kolch, W. Negative regulation of Raf-1 by phosphorylation of serine 621. Mol. Cell. Biol. 1996, 16, 5409–5418. [Google Scholar] [CrossRef] [Green Version]
  86. Diaz, B.; Barnard, D.; Filson, A.; MacDonald, S.; King, A.; Marshall, M. Phosphorylation of Raf-1 serine 338-serine 339 is an essential regulatory event for Ras-dependent activation and biological signaling. Mol. Cell. Biol. 1997, 17, 4509–4516. [Google Scholar] [CrossRef] [Green Version]
  87. Mason, C.S.; Springer, C.J.; Cooper, R.G.; Superti-Furga, G.; Marshall, C.J.; Marais, R. Serine and tyrosine phosphorylations cooperate in Raf-1, but not B-Raf activation. EMBO J. 1999, 18, 2137–2148. [Google Scholar] [CrossRef] [Green Version]
  88. Galabova-Kovacs, G.; Kolbus, A.; Matzen, D.; Meissl, K.; Piazzolla, D.; Rubiolo, C.; Steinitz, K.; Baccarini, M. ERK and beyond: Insights from B-Raf and Raf-1 conditional knockouts. Cell Cycle 2006, 5, 1514–1518. [Google Scholar] [CrossRef]
  89. Sun, H.; King, A.J.; Diaz, H.B.; Marshall, M.S. Regulation of the protein kinase Raf-1 by oncogenic Ras through phosphatidylinositol 3-kinase, Cdc42/Rac and Pak. Curr. Biol. 2000, 10, 281–284. [Google Scholar] [CrossRef] [Green Version]
  90. King, A.J.; Sun, H.; Diaz, B.; Barnard, D.; Miao, W.; Bagrodia, S.; Marshall, M.S. The protein kinase Pak3 positively regulates Raf-1 activity through phosphorylation of serine 338. Nature 1998, 396, 180–183. [Google Scholar] [CrossRef]
  91. Chaudhary, A.; King, W.G.; Mattaliano, M.D.; Frost, J.A.; Diaz, B.; Morrison, D.K.; Cobb, M.H.; Marshall, M.S.; Brugge, J.S. Phosphatidylinositol 3-kinase regulates Raf1 through Pak phosphorylation of serine 338. Curr. Biol. 2000, 10, 551–554. [Google Scholar] [CrossRef] [Green Version]
  92. Zang, M.; Waelde, C.A.; Xiang, X.; Rana, A.; Wen, R.; Luo, Z. Microtubule integrity regulates Pak leading to Ras-independent activation of Raf-1. insights into mechanisms of Raf-1 activation. J. Biol. Chem. 2001, 276, 25157–25165. [Google Scholar] [CrossRef] [Green Version]
  93. Zang, M.; Hayne, C.; Luo, Z. Interaction between active Pak1 and Raf-1 is necessary for phosphorylation and activation of Raf-1. J. Biol. Chem. 2002, 277, 4395–4405. [Google Scholar] [CrossRef] [Green Version]
  94. Zhu, J.; Balan, V.; Bronisz, A.; Balan, K.; Sun, H.; Leicht, D.T.; Luo, Z.; Qin, J.; Avruch, J.; Tzivion, G. Identification of Raf-1 S471 as a novel phosphorylation site critical for Raf-1 and B-Raf kinase activities and for MEK binding. Mol. Biol. Cell 2005, 16, 4733–4744. [Google Scholar] [CrossRef] [Green Version]
  95. Zhang, B.H.; Guan, K.L. Activation of B-Raf kinase requires phosphorylation of the conserved residues Thr598 and Ser601. EMBO J. 2000, 19, 5429–5439. [Google Scholar] [CrossRef] [Green Version]
  96. Carroll, M.P.; May, W.S. Protein kinase C-mediated serine phosphorylation directly activates Raf-1 in murine hematopoietic cells. J. Biol. Chem. 1994, 269, 1249–1256. [Google Scholar] [CrossRef]
  97. Kolch, W.; Heidecker, G.; Kochs, G.; Hummel, R.; Vahidi, H.; Mischak, H.; Finkenzeller, G.; Marmé, D.; Rapp, U.R. Protein kinase C alpha activates RAF-1 by direct phosphorylation. Nature 1993, 364, 249–252. [Google Scholar] [CrossRef]
  98. Cook, S.J.; McCormick, F. Inhibition by cAMP of Ras-dependent activation of Raf. Science 1993, 262, 1069–1072. [Google Scholar] [CrossRef]
  99. Wu, J.; Dent, P.; Jelinek, T.; Wolfman, A.; Weber, M.J.; Sturgill, T.W. Inhibition of the EGF-activated MAP kinase signaling pathway by adenosine 3’,5’-monophosphate. Science 1993, 262, 1065–1069. [Google Scholar] [CrossRef]
  100. Schramm, K.; Niehof, M.; Radziwill, G.; Rommel, C.; Moelling, K. Phosphorylation of c-Raf-1 by protein kinase A interferes with activation. Biochem. Biophys. Res. Commun. 1994, 201, 740–747. [Google Scholar] [CrossRef]
  101. Häfner, S.; Adler, H.S.; Mischak, H.; Janosch, P.; Heidecker, G.; Wolfman, A.; Pippig, S.; Lohse, M.; Ueffing, M.; Kolch, W. Mechanism of inhibition of Raf-1 by protein kinase A. Mol. Cell. Biol. 1994, 14, 6696–6703. [Google Scholar] [CrossRef]
  102. Sidovar, M.F.; Kozlowski, P.; Lee, J.W.; Collins, M.A.; He, Y.; Graves, L.M. Phosphorylation of serine 43 is not required for inhibition of c-Raf kinase by the cAMP-dependent protein kinase. J. Biol. Chem. 2000, 275, 28688–28694. [Google Scholar] [CrossRef] [Green Version]
  103. Dougherty, M.K.; Müller, J.; Ritt, D.A.; Zhou, M.; Zhou, X.Z.; Copeland, T.D.; Conrads, T.P.; Veenstra, T.D.; Lu, K.P.; Morrison, D.K. Regulation of Raf-1 by direct feedback phosphorylation. Mol. Cell 2005, 17, 215–224. [Google Scholar] [CrossRef]
  104. Balan, V.; Leicht, D.T.; Zhu, J.; Balan, K.; Kaplun, A.; Singh-Gupta, V.; Qin, J.; Ruan, H.; Comb, M.J.; Tzivion, G. Identification of novel in vivo Raf-1 phosphorylation sites mediating positive feedback Raf-1 regulation by extracellular signal-regulated kinase. Mol. Biol. Cell 2006, 17, 1141–1153. [Google Scholar] [CrossRef] [Green Version]
  105. Morrison, D.K.; Heidecker, G.; Rapp, U.R.; Copeland, T.D. Identification of the major phosphorylation sites of the Raf-1 kinase. J. Biol. Chem. 1993, 268, 17309–17316. [Google Scholar] [CrossRef]
  106. Dhillon, A.S.; Pollock, C.; Steen, H.; Shaw, P.E.; Mischak, H.; Kolch, W. Cyclic AMP-dependent kinase regulates Raf-1 kinase mainly by phosphorylation of serine 259. Mol. Cell. Biol. 2002, 22, 3237–3246. [Google Scholar] [CrossRef] [Green Version]
  107. Michaud, N.R.; Fabian, J.R.; Mathes, K.D.; Morrison, D.K. 14-3-3 is not essential for Raf-1 function: Identification of Raf-1 proteins that are biologically activated in a 14-3-3- and Ras-independent manner. Mol. Cell. Biol. 1995, 15, 3390–3397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Muslin, A.J.; Tanner, J.W.; Allen, P.M.; Shaw, A.S. Interaction of 14-3-3 with signaling proteins is mediated by the recognition of phosphoserine. Cell 1996, 84, 889–897. [Google Scholar] [CrossRef] [Green Version]
  109. Rommel, C.; Clarke, B.A.; Zimmermann, S.; Nuñez, L.; Rossman, R.; Reid, K.; Moelling, K.; Yancopoulos, G.D.; Glass, D.J. Differentiation stage-specific inhibition of the Raf-MEK-ERK pathway by Akt. Science 1999, 286, 1738–1741. [Google Scholar] [CrossRef]
  110. Zimmermann, S.; Moelling, K. Phosphorylation and regulation of Raf by Akt (protein kinase B). Science 1999, 286, 1741–1744. [Google Scholar] [CrossRef]
  111. Yao, B.; Zhang, Y.; Delikat, S.; Mathias, S.; Basu, S.; Kolesnick, R. Phosphorylation of Raf by ceramide-activated protein kinase. Nature 1995, 378, 307–310. [Google Scholar] [CrossRef] [PubMed]
  112. Zhang, Y.; Yao, B.; Delikat, S.; Bayoumy, S.; Lin, X.H.; Basu, S.; McGinley, M.; Chan-Hui, P.Y.; Lichenstein, H.; Kolesnick, R. Kinase suppressor of Ras is ceramide-activated protein kinase. Cell 1997, 89, 63–72. [Google Scholar] [CrossRef] [Green Version]
  113. Cai, H.; Smola, U.; Wixler, V.; Eisenmann-Tappe, I.; Diaz-Meco, M.T.; Moscat, J.; Rapp, U.; Cooper, G.M. Role of diacylglycerol-regulated protein kinase C isotypes in growth factor activation of the Raf-1 protein kinase. Mol. Cell. Biol. 1997, 17, 732–741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Takahashi, M.; Li, Y.; Dillon, T.J.; Kariya, Y.; Stork, P.J.S. Phosphorylation of the C-Raf N Region Promotes Raf Dimerization. Mol. Cell. Biol. 2017, 37, e00132-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Leicht, D.T.; Balan, V.; Zhu, J.; Kaplun, A.; Bronisz, A.; Rana, A.; Tzivion, G. MEK-1 activates C-Raf through a Ras-independent mechanism. Biochim. Biophys. Acta 2013, 1833, 976–986. [Google Scholar] [CrossRef] [Green Version]
  116. Unni, A.M.; Harbourne, B.; Oh, M.H.; Wild, S.; Ferrarone, J.R.; Lockwood, W.W.; Varmus, H. Hyperactivation of ERK by multiple mechanisms is toxic to RTK-RAS mutation-driven lung adenocarcinoma cells. Elife 2018, 7, e33718. [Google Scholar] [CrossRef]
  117. Leung, G.P.; Feng, T.; Sigoillot, F.D.; Geyer, F.C.; Shirley, M.D.; Ruddy, D.A.; Rakiec, D.P.; Freeman, A.K.; Engelman, J.A.; Jaskelioff, M.; et al. Hyperactivation of MAPK Signaling Is Deleterious to RAS/RAF-mutant Melanoma. Mol. Cancer Res. 2019, 17, 199–211. [Google Scholar] [CrossRef] [Green Version]
  118. Sale, M.J.; Balmanno, K.; Saxena, J.; Ozono, E.; Wojdyla, K.; McIntyre, R.E.; Gilley, R.; Woroniuk, A.; Howarth, K.D.; Hughes, G.; et al. MEK1/2 inhibitor withdrawal reverses acquired resistance driven by BRAF(V600E) amplification whereas KRAS(G13D) amplification promotes EMT-chemoresistance. Nat. Commun. 2019, 10, 2030. [Google Scholar] [CrossRef] [Green Version]
  119. Yeung, K.; Janosch, P.; McFerran, B.; Rose, D.W.; Mischak, H.; Sedivy, J.M.; Kolch, W. Mechanism of suppression of the Raf/MEK/extracellular signal-regulated kinase pathway by the raf kinase inhibitor protein. Mol. Cell. Biol. 2000, 20, 3079–3085. [Google Scholar] [CrossRef] [Green Version]
  120. Yeung, K.; Seitz, T.; Li, S.; Janosch, P.; McFerran, B.; Kaiser, C.; Fee, F.; Katsanakis, K.D.; Rose, D.W.; Mischak, H.; et al. Suppression of Raf-1 kinase activity and MAP kinase signalling by RKIP. Nature 1999, 401, 173–177. [Google Scholar] [CrossRef] [Green Version]
  121. Wakioka, T.; Sasaki, A.; Kato, R.; Shouda, T.; Matsumoto, A.; Miyoshi, K.; Tsuneoka, M.; Komiya, S.; Baron, R.; Yoshimura, A. Spred is a Sprouty-related suppressor of Ras signalling. Nature 2001, 412, 647–651. [Google Scholar] [CrossRef]
  122. Mason, J.M.; Morrison, D.J.; Basson, M.A.; Licht, J.D. Sprouty proteins: Multifaceted negative-feedback regulators of receptor tyrosine kinase signaling. Trends Cell Biol. 2006, 16, 45–54. [Google Scholar] [CrossRef]
  123. Leicht, D.T.; Balan, V.; Kaplun, A.; Singh-Gupta, V.; Kaplun, L.; Dobson, M.; Tzivion, G. Raf kinases: Function, regulation and role in human cancer. Biochim. Biophys. Acta 2007, 1773, 1196–1212. [Google Scholar] [CrossRef] [Green Version]
  124. Wellbrock, C.; Karasarides, M.; Marais, R. The RAF proteins take centre stage. Nat. Rev. Mol. Cell. Biol. 2004, 5, 875–885. [Google Scholar] [CrossRef]
  125. Dhillon, A.S.; Hagan, S.; Rath, O.; Kolch, W. MAP kinase signalling pathways in cancer. Oncogene 2007, 26, 3279–3290. [Google Scholar] [CrossRef] [Green Version]
  126. Wu, X.; Noh, S.J.; Zhou, G.; Dixon, J.E.; Guan, K.L. Selective activation of MEK1 but not MEK2 by A-Raf from epidermal growth factor-stimulated Hela cells. J. Biol. Chem. 1996, 271, 3265–3271. [Google Scholar] [CrossRef] [Green Version]
  127. Frost, J.A.; Steen, H.; Shapiro, P.; Lewis, T.; Ahn, N.; Shaw, P.E.; Cobb, M.H. Cross-cascade activation of ERKs and ternary complex factors by Rho family proteins. EMBO J. 1997, 16, 6426–6438. [Google Scholar] [CrossRef]
  128. Coles, L.C.; Shaw, P.E. PAK1 primes MEK1 for phosphorylation by Raf-1 kinase during cross-cascade activation of the ERK pathway. Oncogene 2002, 21, 2236–2244. [Google Scholar] [CrossRef] [Green Version]
  129. Kholodenko, B.N.; Hancock, J.F.; Kolch, W. Signalling ballet in space and time. Nat. Rev. Mol. Cell. Biol. 2010, 11, 414–426. [Google Scholar] [CrossRef] [Green Version]
  130. Kornfeld, K.; Hom, D.B.; Horvitz, H.R. The ksr-1 gene encodes a novel protein kinase involved in Ras-mediated signaling in C. elegans. Cell 1995, 83, 903–913. [Google Scholar] [CrossRef] [Green Version]
  131. Sundaram, M.; Han, M. The C. elegans ksr-1 gene encodes a novel Raf-related kinase involved in Ras-mediated signal transduction. Cell 1995, 83, 889–901. [Google Scholar] [CrossRef] [Green Version]
  132. Therrien, M.; Chang, H.C.; Solomon, N.M.; Karim, F.D.; Wassarman, D.A.; Rubin, G.M. KSR, a novel protein kinase required for RAS signal transduction. Cell 1995, 83, 879–888. [Google Scholar] [CrossRef] [Green Version]
  133. Therrien, M.; Wong, A.M.; Rubin, G.M. CNK, a RAF-binding multidomain protein required for RAS signaling. Cell 1998, 95, 343–353. [Google Scholar] [CrossRef] [Green Version]
  134. Sieburth, D.S.; Sun, Q.; Han, M. SUR-8, a conserved Ras-binding protein with leucine-rich repeats, positively regulates Ras-mediated signaling in C. elegans. Cell 1998, 94, 119–130. [Google Scholar] [CrossRef] [Green Version]
  135. Luttrell, L.M.; Roudabush, F.L.; Choy, E.W.; Miller, W.E.; Field, M.E.; Pierce, K.L.; Lefkowitz, R.J. Activation and targeting of extracellular signal-regulated kinases by beta-arrestin scaffolds. Proc. Natl. Acad. Sci. USA 2001, 98, 2449–2454. [Google Scholar] [CrossRef] [Green Version]
  136. Ishibe, S.; Joly, D.; Liu, Z.X.; Cantley, L.G. Paxillin serves as an ERK-regulated scaffold for coordinating FAK and Rac activation in epithelial morphogenesis. Mol. Cell 2004, 16, 257–267. [Google Scholar] [CrossRef]
  137. Sharma, C.; Vomastek, T.; Tarcsafalvi, A.; Catling, A.D.; Schaeffer, H.J.; Eblen, S.T.; Weber, M.J. MEK partner 1 (MP1): Regulation of oligomerization in MAP kinase signaling. J. Cell Biochem. 2005, 94, 708–719. [Google Scholar] [CrossRef] [PubMed]
  138. McKay, M.M.; Ritt, D.A.; Morrison, D.K. Signaling dynamics of the KSR1 scaffold complex. Proc. Natl. Acad. Sci. USA 2009, 106, 11022–11027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Rajakulendran, T.; Sahmi, M.; Lefrançois, M.; Sicheri, F.; Therrien, M. A dimerization-dependent mechanism drives RAF catalytic activation. Nature 2009, 461, 542–545. [Google Scholar] [CrossRef] [PubMed]
  140. Nguyen, A.; Burack, W.R.; Stock, J.L.; Kortum, R.; Chaika, O.V.; Afkarian, M.; Muller, W.J.; Murphy, K.M.; Morrison, D.K.; Lewis, R.E.; et al. Kinase suppressor of Ras (KSR) is a scaffold which facilitates mitogen-activated protein kinase activation in vivo. Mol. Cell. Biol. 2002, 22, 3035–3045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Smith, F.D.; Langeberg, L.K.; Cellurale, C.; Pawson, T.; Morrison, D.K.; Davis, R.J.; Scott, J.D. AKAP-Lbc enhances cyclic AMP control of the ERK1/2 cascade. Nat. Cell Biol. 2010, 12, 1242–1249. [Google Scholar] [CrossRef]
  142. Dougherty, M.K.; Ritt, D.A.; Zhou, M.; Specht, S.I.; Monson, D.M.; Veenstra, T.D.; Morrison, D.K. KSR2 is a calcineurin substrate that promotes ERK cascade activation in response to calcium signals. Mol. Cell 2009, 34, 652–662. [Google Scholar] [CrossRef] [Green Version]
  143. Pritchard, C.A.; Bolin, L.; Slattery, R.; Murray, R.; McMahon, M. Post-natal lethality and neurological and gastrointestinal defects in mice with targeted disruption of the A-Raf protein kinase gene. Curr. Biol. 1996, 6, 614–617. [Google Scholar] [CrossRef] [Green Version]
  144. Wojnowski, L.; Zimmer, A.M.; Beck, T.W.; Hahn, H.; Bernal, R.; Rapp, U.R.; Zimmer, A. Endothelial apoptosis in Braf-deficient mice. Nat. Genet. 1997, 16, 293–297. [Google Scholar] [CrossRef]
  145. Wojnowski, L.; Stancato, L.F.; Zimmer, A.M.; Hahn, H.; Beck, T.W.; Larner, A.C.; Rapp, U.R.; Zimmer, A. Craf-1 protein kinase is essential for mouse development. Mech. Dev. 1998, 76, 141–149. [Google Scholar] [CrossRef]
  146. Wiese, S.; Pei, G.; Karch, C.; Troppmair, J.; Holtmann, B.; Rapp, U.R.; Sendtner, M. Specific function of B-Raf in mediating survival of embryonic motoneurons and sensory neurons. Nat. Neurosci. 2001, 4, 137–142. [Google Scholar] [CrossRef]
  147. De Iriarte Rodríguez, R.; Magariños, M.; Pfeiffer, V.; Rapp, U.R.; Varela-Nieto, I. C-Raf deficiency leads to hearing loss and increased noise susceptibility. Cell Mol. Life Sci. 2015, 72, 3983–3998. [Google Scholar] [CrossRef] [Green Version]
  148. Hüser, M.; Luckett, J.; Chiloeches, A.; Mercer, K.; Iwobi, M.; Giblett, S.; Sun, X.M.; Brown, J.; Marais, R.; Pritchard, C. MEK kinase activity is not necessary for Raf-1 function. EMBO J. 2001, 20, 1940–1951. [Google Scholar] [CrossRef] [Green Version]
  149. Mikula, M.; Schreiber, M.; Husak, Z.; Kucerova, L.; Rüth, J.; Wieser, R.; Zatloukal, K.; Beug, H.; Wagner, E.F.; Baccarini, M. Embryonic lethality and fetal liver apoptosis in mice lacking the c-raf-1 gene. EMBO J. 2001, 20, 1952–1962. [Google Scholar] [CrossRef] [Green Version]
  150. Rebocho, A.P.; Marais, R. ARAF acts as a scaffold to stabilize BRAF:CRAF heterodimers. Oncogene 2013, 32, 3207–3212. [Google Scholar] [CrossRef] [Green Version]
  151. Dumaz, N.; Hayward, R.; Martin, J.; Ogilvie, L.; Hedley, D.; Curtin, J.A.; Bastian, B.C.; Springer, C.; Marais, R. In melanoma, RAS mutations are accompanied by switching signaling from BRAF to CRAF and disrupted cyclic AMP signaling. Cancer Res. 2006, 66, 9483–9491. [Google Scholar] [CrossRef] [Green Version]
  152. McPhillips, F.; Mullen, P.; MacLeod, K.G.; Sewell, J.M.; Monia, B.P.; Cameron, D.A.; Smyth, J.F.; Langdon, S.P. Raf-1 is the predominant Raf isoform that mediates growth factor-stimulated growth in ovarian cancer cells. Carcinogenesis 2006, 27, 729–739. [Google Scholar] [CrossRef] [Green Version]
  153. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef]
  154. Wan, P.T.; Garnett, M.J.; Roe, S.M.; Lee, S.; Niculescu-Duvaz, D.; Good, V.M.; Jones, C.M.; Marshall, C.J.; Springer, C.J.; Barford, D.; et al. Mechanism of activation of the RAF-ERK signaling pathway by oncogenic mutations of B-RAF. Cell 2004, 116, 855–867. [Google Scholar] [CrossRef] [Green Version]
  155. Kumar, R.; Angelini, S.; Czene, K.; Sauroja, I.; Hahka-Kemppinen, M.; Pyrhönen, S.; Hemminki, K. BRAF mutations in metastatic melanoma: A possible association with clinical outcome. Clin. Cancer Res. 2003, 9, 3362–3368. [Google Scholar]
  156. Dhomen, N.; Marais, R. New insight into BRAF mutations in cancer. Curr. Opin. Genet. Dev. 2007, 17, 31–39. [Google Scholar] [CrossRef]
  157. Garnett, M.J.; Rana, S.; Paterson, H.; Barford, D.; Marais, R. Wild-type and mutant B-RAF activate C-RAF through distinct mechanisms involving heterodimerization. Mol. Cell 2005, 20, 963–969. [Google Scholar] [CrossRef]
  158. Tran, N.H.; Frost, J.A. Phosphorylation of Raf-1 by p21-activated kinase 1 and Src regulates Raf-1 autoinhibition. J. Biol. Chem. 2003, 278, 11221–11226. [Google Scholar] [CrossRef] [Green Version]
  159. Zebisch, A.; Staber, P.B.; Delavar, A.; Bodner, C.; Hiden, K.; Fischereder, K.; Janakiraman, M.; Linkesch, W.; Auner, H.W.; Emberger, W.; et al. Two transforming C-RAF germ-line mutations identified in patients with therapy-related acute myeloid leukemia. Cancer Res. 2006, 66, 3401–3408. [Google Scholar] [CrossRef] [Green Version]
  160. Pandit, B.; Sarkozy, A.; Pennacchio, L.A.; Carta, C.; Oishi, K.; Martinelli, S.; Pogna, E.A.; Schackwitz, W.; Ustaszewska, A.; Landstrom, A.; et al. Gain-of-function RAF1 mutations cause Noonan and LEOPARD syndromes with hypertrophic cardiomyopathy. Nat. Genet. 2007, 39, 1007–1012. [Google Scholar] [CrossRef]
  161. Razzaque, M.A.; Nishizawa, T.; Komoike, Y.; Yagi, H.; Furutani, M.; Amo, R.; Kamisago, M.; Momma, K.; Katayama, H.; Nakagawa, M.; et al. Germline gain-of-function mutations in RAF1 cause Noonan syndrome. Nat. Genet. 2007, 39, 1013–1017. [Google Scholar] [CrossRef] [PubMed]
  162. Rauen, K.A. The RASopathies. Annu. Rev. Genomics Hum. Genet. 2013, 14, 355–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Yao, Z.; Torres, N.M.; Tao, A.; Gao, Y.; Luo, L.; Li, Q.; de Stanchina, E.; Abdel-Wahab, O.; Solit, D.B.; Poulikakos, P.I.; et al. BRAF Mutants Evade ERK-Dependent Feedback by Different Mechanisms that Determine Their Sensitivity to Pharmacologic Inhibition. Cancer Cell 2015, 28, 370–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Yao, Z.; Yaeger, R.; Rodrik-Outmezguine, V.S.; Tao, A.; Torres, N.M.; Chang, M.T.; Drosten, M.; Zhao, H.; Cecchi, F.; Hembrough, T.; et al. Tumours with class 3 BRAF mutants are sensitive to the inhibition of activated RAS. Nature 2017, 548, 234–238. [Google Scholar] [CrossRef]
  165. Yaeger, R.; Corcoran, R.B. Targeting Alterations in the RAF-MEK Pathway. Cancer Discov. 2019, 9, 329–341. [Google Scholar] [CrossRef] [Green Version]
  166. Lavoie, H.; Thevakumaran, N.; Gavory, G.; Li, J.J.; Padeganeh, A.; Guiral, S.; Duchaine, J.; Mao, D.Y.; Bouvier, M.; Sicheri, F.; et al. Inhibitors that stabilize a closed RAF kinase domain conformation induce dimerization. Nat. Chem. Biol. 2013, 9, 428–436. [Google Scholar] [CrossRef]
  167. Shaw, A.S.; Kornev, A.P.; Hu, J.; Ahuja, L.G.; Taylor, S.S. Kinases and pseudokinases: Lessons from RAF. Mol. Cell. Biol. 2014, 34, 1538–1546. [Google Scholar] [CrossRef] [Green Version]
  168. Karoulia, Z.; Wu, Y.; Ahmed, T.A.; Xin, Q.; Bollard, J.; Krepler, C.; Wu, X.; Zhang, C.; Bollag, G.; Herlyn, M.; et al. An Integrated Model of RAF Inhibitor Action Predicts Inhibitor Activity against Oncogenic BRAF Signaling. Cancer Cell 2016, 30, 485–498. [Google Scholar] [CrossRef] [Green Version]
  169. Lyons, J.F.; Wilhelm, S.; Hibner, B.; Bollag, G. Discovery of a novel Raf kinase inhibitor. Endocr. Relat. Cancer 2001, 8, 219–225. [Google Scholar] [CrossRef] [Green Version]
  170. Poulikakos, P.I.; Zhang, C.; Bollag, G.; Shokat, K.M.; Rosen, N. RAF inhibitors transactivate RAF dimers and ERK signalling in cells with wild-type BRAF. Nature 2010, 464, 427–430. [Google Scholar] [CrossRef] [Green Version]
  171. Hall-Jackson, C.A.; Eyers, P.A.; Cohen, P.; Goedert, M.; Boyle, F.T.; Hewitt, N.; Plant, H.; Hedge, P. Paradoxical activation of Raf by a novel Raf inhibitor. Chem. Biol. 1999, 6, 559–568. [Google Scholar] [CrossRef]
  172. Bollag, G.; Hirth, P.; Tsai, J.; Zhang, J.; Ibrahim, P.N.; Cho, H.; Spevak, W.; Zhang, C.; Zhang, Y.; Habets, G.; et al. Clinical efficacy of a RAF inhibitor needs broad target blockade in BRAF-mutant melanoma. Nature 2010, 467, 596–599. [Google Scholar] [CrossRef]
  173. Hauschild, A.; Grob, J.J.; Demidov, L.V.; Jouary, T.; Gutzmer, R.; Millward, M.; Rutkowski, P.; Blank, C.U.; Miller, W.H., Jr.; Kaempgen, E.; et al. Dabrafenib in BRAF-mutated metastatic melanoma: A multicentre, open-label, phase 3 randomised controlled trial. Lancet 2012, 380, 358–365. [Google Scholar] [CrossRef]
  174. Chapman, P.B.; Hauschild, A.; Robert, C.; Haanen, J.B.; Ascierto, P.; Larkin, J.; Dummer, R.; Garbe, C.; Testori, A.; Maio, M.; et al. Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N. Engl. J. Med. 2011, 364, 2507–2516. [Google Scholar] [CrossRef] [Green Version]
  175. Lito, P.; Pratilas, C.A.; Joseph, E.W.; Tadi, M.; Halilovic, E.; Zubrowski, M.; Huang, A.; Wong, W.L.; Callahan, M.K.; Merghoub, T.; et al. Relief of profound feedback inhibition of mitogenic signaling by RAF inhibitors attenuates their activity in BRAFV600E melanomas. Cancer Cell 2012, 22, 668–682. [Google Scholar] [CrossRef] [Green Version]
  176. Moriceau, G.; Hugo, W.; Hong, A.; Shi, H.; Kong, X.; Yu, C.C.; Koya, R.C.; Samatar, A.A.; Khanlou, N.; Braun, J.; et al. Tunable-combinatorial mechanisms of acquired resistance limit the efficacy of BRAF/MEK cotargeting but result in melanoma drug addiction. Cancer Cell 2015, 27, 240–256. [Google Scholar] [CrossRef] [Green Version]
  177. Zhang, C.; Spevak, W.; Zhang, Y.; Burton, E.A.; Ma, Y.; Habets, G.; Zhang, J.; Lin, J.; Ewing, T.; Matusow, B.; et al. RAF inhibitors that evade paradoxical MAPK pathway activation. Nature 2015, 526, 583–586. [Google Scholar] [CrossRef]
  178. Hatzivassiliou, G.; Song, K.; Yen, I.; Brandhuber, B.J.; Anderson, D.J.; Alvarado, R.; Ludlam, M.J.; Stokoe, D.; Gloor, S.L.; Vigers, G.; et al. RAF inhibitors prime wild-type RAF to activate the MAPK pathway and enhance growth. Nature 2010, 464, 431–435. [Google Scholar] [CrossRef] [Green Version]
  179. Man, R.J.; Zhang, Y.L.; Jiang, A.Q.; Zhu, H.L. A patent review of RAF kinase inhibitors (2010–2018). Expert Opin. Ther. Pat. 2019, 29, 675–688. [Google Scholar] [CrossRef]
  180. Henry, J.R.; Kaufman, M.D.; Peng, S.B.; Ahn, Y.M.; Caldwell, T.M.; Vogeti, L.; Telikepalli, H.; Lu, W.P.; Hood, M.M.; Rutkoski, T.J.; et al. Discovery of 1-(3,3-dimethylbutyl)-3-(2-fluoro-4-methyl-5-(7-methyl-2-(methylamino)pyrido [2,3-d]pyrimidin-6-yl)phenyl)urea (LY3009120) as a pan-RAF inhibitor with minimal paradoxical activation and activity against BRAF or RAS mutant tumor cells. J. Med. Chem. 2015, 58, 4165–4179. [Google Scholar] [CrossRef]
  181. Peng, S.B.; Henry, J.R.; Kaufman, M.D.; Lu, W.P.; Smith, B.D.; Vogeti, S.; Rutkoski, T.J.; Wise, S.; Chun, L.; Zhang, Y.; et al. Inhibition of RAF Isoforms and Active Dimers by LY3009120 Leads to Anti-tumor Activities in RAS or BRAF Mutant Cancers. Cancer Cell 2015, 28, 384–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Noeparast, A.; Giron, P.; De Brakeleer, S.; Eggermont, C.; De Ridder, U.; Teugels, E.; De Grève, J. Type II RAF inhibitor causes superior ERK pathway suppression compared to type I RAF inhibitor in cells expressing different BRAF mutant types recurrently found in lung cancer. Oncotarget 2018, 9, 16110–16123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Girotti, M.R.; Lopes, F.; Preece, N.; Niculescu-Duvaz, D.; Zambon, A.; Davies, L.; Whittaker, S.; Saturno, G.; Viros, A.; Pedersen, M.; et al. Paradox-breaking RAF inhibitors that also target SRC are effective in drug-resistant BRAF mutant melanoma. Cancer Cell 2015, 27, 85–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Sun, Y.; Alberta, J.A.; Pilarz, C.; Calligaris, D.; Chadwick, E.J.; Ramkissoon, S.H.; Ramkissoon, L.A.; Garcia, V.M.; Mazzola, E.; Goumnerova, L.; et al. A brain-penetrant RAF dimer antagonist for the noncanonical BRAF oncoprotein of pediatric low-grade astrocytomas. Neuro Oncol. 2017, 19, 774–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Sullivan, R.J.; Hollebecque, A.; Flaherty, K.T.; Shapiro, G.I.; Rodon Ahnert, J.; Millward, M.J.; Zhang, W.; Gao, L.; Sykes, A.; Willard, M.D.; et al. A Phase I Study of LY3009120, a Pan-RAF Inhibitor, in Patients with Advanced or Metastatic Cancer. Mol. Cancer Ther. 2020, 19, 460–467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Koumaki, K.; Kontogianni, G.; Kosmidou, V.; Pahitsa, F.; Kritsi, E.; Zervou, M.; Chatziioannou, A.; Souliotis, V.L.; Papadodima, O.; Pintzas, A. BRAF paradox breakers PLX8394, PLX7904 are more effective against BRAFV600Ε CRC cells compared with the BRAF inhibitor PLX4720 and shown by detailed pathway analysis. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166061. [Google Scholar] [CrossRef]
  187. Yao, Z.; Gao, Y.; Su, W.; Yaeger, R.; Tao, J.; Na, N.; Zhang, Y.; Zhang, C.; Rymar, A.; Tao, A.; et al. RAF inhibitor PLX8394 selectively disrupts BRAF dimers and RAS-independent BRAF-mutant-driven signaling. Nat. Med. 2019, 25, 284–291. [Google Scholar] [CrossRef]
  188. Dumaz, N.; Lebbé, C. New perspectives on targeting RAF, MEK and ERK in melanoma. Curr. Opin. Oncol. 2021, 33, 120–126. [Google Scholar] [CrossRef]
  189. Turajlic, S.; Ali, Z.; Yousaf, N.; Larkin, J. Phase I/II RAF kinase inhibitors in cancer therapy. Expert Opin. Investig. Drugs 2013, 22, 739–749. [Google Scholar] [CrossRef]
  190. Basile, K.J.; Le, K.; Hartsough, E.J.; Aplin, A.E. Inhibition of mutant BRAF splice variant signaling by next-generation, selective RAF inhibitors. Pigment Cell Melanoma Res. 2014, 27, 479–484. [Google Scholar] [CrossRef]
  191. Long, G.V.; Stroyakovskiy, D.; Gogas, H.; Levchenko, E.; de Braud, F.; Larkin, J.; Garbe, C.; Jouary, T.; Hauschild, A.; Grob, J.J.; et al. Combined BRAF and MEK inhibition versus BRAF inhibition alone in melanoma. N. Engl. J. Med. 2014, 371, 1877–1888. [Google Scholar] [CrossRef] [Green Version]
  192. Larkin, J.; Ascierto, P.A.; Dréno, B.; Atkinson, V.; Liszkay, G.; Maio, M.; Mandalà, M.; Demidov, L.; Stroyakovskiy, D.; Thomas, L.; et al. Combined vemurafenib and cobimetinib in BRAF-mutated melanoma. N. Engl. J. Med. 2014, 371, 1867–1876. [Google Scholar] [CrossRef] [Green Version]
  193. Dummer, R.; Hauschild, A.; Santinami, M.; Atkinson, V.; Mandalà, M.; Kirkwood, J.M.; Chiarion Sileni, V.; Larkin, J.; Nyakas, M.; Dutriaux, C.; et al. Five-Year Analysis of Adjuvant Dabrafenib plus Trametinib in Stage III Melanoma. N. Engl. J. Med. 2020, 383, 1139–1148. [Google Scholar] [CrossRef]
  194. Long, G.V.; Hauschild, A.; Santinami, M.; Atkinson, V.; Mandalà, M.; Chiarion-Sileni, V.; Larkin, J.; Nyakas, M.; Dutriaux, C.; Haydon, A.; et al. Adjuvant Dabrafenib plus Trametinib in Stage III BRAF-Mutated Melanoma. N. Engl. J. Med. 2017, 377, 1813–1823. [Google Scholar] [CrossRef] [Green Version]
  195. Gogas, H.; Dréno, B.; Larkin, J.; Demidov, L.; Stroyakovskiy, D.; Eroglu, Z.; Francesco Ferrucci, P.; Pigozzo, J.; Rutkowski, P.; Mackiewicz, J.; et al. Cobimetinib plus atezolizumab in BRAF(V600) wild-type melanoma: Primary results from the randomized phase III IMspire170 study. Ann. Oncol. 2021, 32, 384–394. [Google Scholar] [CrossRef]
  196. Dummer, R.; Ascierto, P.A.; Gogas, H.J.; Arance, A.; Mandala, M.; Liszkay, G.; Garbe, C.; Schadendorf, D.; Krajsova, I.; Gutzmer, R.; et al. Encorafenib plus binimetinib versus vemurafenib or encorafenib in patients with BRAF-mutant melanoma (COLUMBUS): A multicentre, open-label, randomised phase 3 trial. Lancet Oncol. 2018, 19, 603–615. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structure of Raf family kinases. All Raf isoforms comprise three conserved regions: conserved region 1 (CR1) contains a Ras-binding domain (RBD) and a Cys-rich domain (CRD); conserved region 2 (CR2) is characterized by a Ser/Thr-rich sequence where 14-3-3 binds and inhibits Raf; conserved region 3 (CR3) is the kinase domain where the B-Raf V600E mutation is found in cancer. At the C-terminus, the second site promotes dimerization via binding to 14-3-3. BRS is a B-Raf-specific site. The viral oncoproteins v-Raf and v-mil have amino-terminal truncations and are fused with the N-myristoylated (N-myr) viral Gag protein. Four conserved phosphorylation sites of each Raf isoform are indicated in rectangles, including 14-3-3 binding sites and phoshorylattion sites in the activation loop. NtA: N-terminal acidic region.
Figure 1. Structure of Raf family kinases. All Raf isoforms comprise three conserved regions: conserved region 1 (CR1) contains a Ras-binding domain (RBD) and a Cys-rich domain (CRD); conserved region 2 (CR2) is characterized by a Ser/Thr-rich sequence where 14-3-3 binds and inhibits Raf; conserved region 3 (CR3) is the kinase domain where the B-Raf V600E mutation is found in cancer. At the C-terminus, the second site promotes dimerization via binding to 14-3-3. BRS is a B-Raf-specific site. The viral oncoproteins v-Raf and v-mil have amino-terminal truncations and are fused with the N-myristoylated (N-myr) viral Gag protein. Four conserved phosphorylation sites of each Raf isoform are indicated in rectangles, including 14-3-3 binding sites and phoshorylattion sites in the activation loop. NtA: N-terminal acidic region.
Ijms 23 05158 g001
Figure 2. The Ras/Raf/MEK/ERK signaling cascade. Growth factor binds to receptor tyrosine kinase and activates growth-factor receptor-bound 2 (GRB2) and Son of Sevenless (SOS) to load GTP to Ras. Then, Ras-GTP recruits Raf to the plasma membrane where Raf is activated, leading to sequential phosphorylation and activation of MEK and ERK. Activated ERK then phosphorylates a variety of substrates and elicits various cellular responses. RBD-CRD-CR3 designates essential domains of Raf and CR3 is the kinase domain.
Figure 2. The Ras/Raf/MEK/ERK signaling cascade. Growth factor binds to receptor tyrosine kinase and activates growth-factor receptor-bound 2 (GRB2) and Son of Sevenless (SOS) to load GTP to Ras. Then, Ras-GTP recruits Raf to the plasma membrane where Raf is activated, leading to sequential phosphorylation and activation of MEK and ERK. Activated ERK then phosphorylates a variety of substrates and elicits various cellular responses. RBD-CRD-CR3 designates essential domains of Raf and CR3 is the kinase domain.
Ijms 23 05158 g002
Figure 3. Functional classes of B-Raf mutations. Class I B-Raf mutants contain V600E/D mutations in the activation loop which can signal as active monomers, independent of Ras. Class II B-Raf mutants are Ras-independent and signal as dimers. Class III B-Raf mutants have reduced kinase activity and drive the activation of ERK signaling by transactivating wild-type Raf which signals as mutant B-Raf–wild-type C-Raf dimers. These mutants require active Ras to trigger a signaling cascade.
Figure 3. Functional classes of B-Raf mutations. Class I B-Raf mutants contain V600E/D mutations in the activation loop which can signal as active monomers, independent of Ras. Class II B-Raf mutants are Ras-independent and signal as dimers. Class III B-Raf mutants have reduced kinase activity and drive the activation of ERK signaling by transactivating wild-type Raf which signals as mutant B-Raf–wild-type C-Raf dimers. These mutants require active Ras to trigger a signaling cascade.
Ijms 23 05158 g003
Figure 4. Conformation transition of the Raf kinase domain. The kinase domain consists of an N-terminal lobe (N-lobe) and a C-terminal lobe (C-lobe) linked through a hinge. αC-helix and DFG (green) alter from OUT to IN position upon Raf activation, resulting in a dimer with side-to-side interface “closed” conformation. The conformational transition between inactive and active states is shown.
Figure 4. Conformation transition of the Raf kinase domain. The kinase domain consists of an N-terminal lobe (N-lobe) and a C-terminal lobe (C-lobe) linked through a hinge. αC-helix and DFG (green) alter from OUT to IN position upon Raf activation, resulting in a dimer with side-to-side interface “closed” conformation. The conformational transition between inactive and active states is shown.
Ijms 23 05158 g004
Figure 5. Functional properties of different RAF inhibitors. The upper part shows the effect of RAF inhibitors on monomeric RAF kinases; the lower part shows the effect of RAF inhibitors on dimeric RAF kinases. The first and second generations of Raf inhibitors lead to paradoxical activation in dimeric kinases. The inhibitor binds to one protomer within an RAF dimer, causing conformational change and decreasing the affinity of the inhibitor for the other protomer, as well as substantial transactivation of this protomer (dotted arrow), resulting in higher downstream signaling activation. The third generations of Pan-Raf inhibitors bind to monomeric and dimeric kinases with similar affinity. The third generation of paradox breakers disrupt the B-Raf dimer interface and specifically inhibit B-Raf dimerization but not C-Raf homodimerization, although they bind to C-Raf. INH: inhibitor. RAF designates any isoforms of Raf.
Figure 5. Functional properties of different RAF inhibitors. The upper part shows the effect of RAF inhibitors on monomeric RAF kinases; the lower part shows the effect of RAF inhibitors on dimeric RAF kinases. The first and second generations of Raf inhibitors lead to paradoxical activation in dimeric kinases. The inhibitor binds to one protomer within an RAF dimer, causing conformational change and decreasing the affinity of the inhibitor for the other protomer, as well as substantial transactivation of this protomer (dotted arrow), resulting in higher downstream signaling activation. The third generations of Pan-Raf inhibitors bind to monomeric and dimeric kinases with similar affinity. The third generation of paradox breakers disrupt the B-Raf dimer interface and specifically inhibit B-Raf dimerization but not C-Raf homodimerization, although they bind to C-Raf. INH: inhibitor. RAF designates any isoforms of Raf.
Ijms 23 05158 g005
Table 2. B-Raf inhibitors in cancer therapy.
Table 2. B-Raf inhibitors in cancer therapy.
RAF InhibitorMechanismClinical StageFeatures
First generation
Sorafenib‘αC-IN’/‘DFG-OUT’ inhibitorApproved for advanced renal cell carcinoma and hepatocellular carcinomaTransactivation of ERK1/2 pathway in WT B-Raf cells
Second generation
Vemurafenib‘αC-OUT’/‘DFG-IN’ inhibitorApproved for B-Raf-V600E metastatic melanomaCauses photosensitivity, development of drug resistance and tumor recurrence
Dabrafenib‘αC-OUT’/‘DFG-IN’ inhibitorApproved for melanoma patients with B-Raf-V600E/K mutationsCauses fever, development of drug resistance and tumor recurrence
Third generation
CCT196969‘αC-IN’/‘DFG-OUT’ inhibitorAntitumor activity in preclinical studies against B-Raf-V600E melanomas, Ras-mutant melanomas and colorectal tumorsDual pan-Raf and SRC kinase inhibitor, effective in patient-derived xenograft (PDX) models that included melanomas with intrinsic or acquired resistance to second-generation Raf and MEK inhibitors
CCT241161‘αC-IN’/‘DFG-OUT’ inhibitorAntitumor activity in preclinical studies against B-Raf-V600E melanomas, Ras-mutant melanomas and colorectal tumorsDual pan-Raf and SRC kinase inhibitor, effective in patient-derived xenograft (PDX) models that included melanomas with intrinsic or acquired resistance to second-generation Raf and MEK inhibitors
LY3009120‘αC-IN’/‘DFG-OUT’ inhibitorAntitumor activity in Phase I clinical studies against NRas or KRas mutant tumors and B-Raf deletions in pancreatic and thyroid tumorsEffective in vemurafenib-resistant melanomas; inhibit monomeric and dimeric B-Raf with similar potency
TAK-580 (MLN2480)‘αC-IN’/‘DFG-OUT’ inhibitorAntiproliferative activity in Phase I clinical studies against melanomas and other solid tumor cell lines harboring B-Raf, NRas or KRas mutations; delay emergence of resistanceEffective in vemurafenib-resistant melanomas harboring B-Raf or N-Ras mutations and B-Raf-V600E colorectal or thyroid tumors
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, J.; Luo, Z. Discovery of Raf Family Is a Milestone in Deciphering the Ras-Mediated Intracellular Signaling Pathway. Int. J. Mol. Sci. 2022, 23, 5158. https://doi.org/10.3390/ijms23095158

AMA Style

Zhao J, Luo Z. Discovery of Raf Family Is a Milestone in Deciphering the Ras-Mediated Intracellular Signaling Pathway. International Journal of Molecular Sciences. 2022; 23(9):5158. https://doi.org/10.3390/ijms23095158

Chicago/Turabian Style

Zhao, Jingtong, and Zhijun Luo. 2022. "Discovery of Raf Family Is a Milestone in Deciphering the Ras-Mediated Intracellular Signaling Pathway" International Journal of Molecular Sciences 23, no. 9: 5158. https://doi.org/10.3390/ijms23095158

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop