Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review MAP Kinase-Interacting Kinases—Emerging Targets against Cancer Sarah Diab,1 Malika Kumarasiri,1 Mingfeng Yu,1 Theodosia Teo,1 Christopher Proud,2 Robert Milne,1 and Shudong Wang1,* 1Centre for Drug Discovery and Development, Sansom Institute for Health Research, School of Pharmacy and Medical Sciences, University of South Australia, Adelaide, SA 5001, Australia 2Centre for Biological Sciences, University of Southampton, Southampton SO17 1BJ, UK *Correspondence: [email protected] http://dx.doi.org/10.1016/j.chembiol.2014.01.011

Mitogen-activated protein kinase (MAPK)-interacting kinases (Mnks) regulate the initiation of translation through phosphorylation of eukaryotic initiation factor 4E (eIF4E). Mnk-mediated eIF4E activation promotes cancer development and progression. While the phosphorylation of eIF4E is necessary for oncogenic transformation, the kinase activity of Mnks seems dispensable for normal development. For this reason, pharmacological inhibition of Mnks could represent an ideal mechanism-based and nontoxic therapeutic strategy for cancer treatment. In this review, we discuss the current understanding of Mnk biological roles, structures, and functions, as well as clinical implications. Importantly, we propose different strategies for identification of highly selective small molecule inhibitors of Mnks, including exploring a structural feature of their kinase domain, DFD motif, which is unique within the human kinome. We also argue that a combined targeting of Mnks and other pathways should be considered given the complexity of cancer. The translational process within the cell has become a ‘‘hot spot’’ for cancer therapy since it became apparent that deregulation of protein synthesis is a common event in cancer. This has stimulated enthusiasm for innovative therapies targeting control over the initiation of translation, especially the eukaryotic initiation factor 4E (eIF4E). Given that the phosphorylation of eIF4E by mitogen-activated protein kinase (MAPK)-interacting kinases (Mnks) seems important for its oncogenic activity, Mnks have emerged as potential anticancer therapeutic targets. Our previous review presented strong biological and pharmacological evidence supporting Mnks as targets (Hou et al., 2012). As a complementary review, we provide herein an updated overview of the roles of Mnks in tumor biology, the rational design of pharmacological inhibitors, and potential drug development strategies. The availability of eIF4E to participate in the formation of the eukaryotic initiation complex 4F (which also consists of a scaffold protein eIF4G and an RNA helicase eIF4A) is critical for initiating translation of most mRNAs (Figure 1). Recognition of the 7-methyl guanosine (m7G) cap structure at the 50 -UTR region of mRNAs by eIF4E facilitates recruitment of the 43S preinitiation complex formed by the small ribosomal unit and various initiation factors (Rhoads, 1988). Once the small ribosomal subunit 40S is bound to the 50 -UTR of the mRNAs, it scans along the 50 -leader of the mRNA to locate the initiation codon (usually AUG). The large ribosomal subunit 60S subsequently binds to form a complete ribosome, which proceeds to the elongation cycle. Given its restricted availability, eIF4E is often considered as a rate-limiting factor for translation initiation in eukaryotic cells. Furthermore, it has recently been demonstrated that eIF4E stimulates eIF4A activity within the eIF4F complex and promotes mRNA restructuring and translation independently of the eIF4E cap-binding mechanism (Feoktistova et al., 2013). This might shed light on the discriminatory effect of eIF4E on the translation

of capped mRNAs: physiologically, a low level of eIF4E is sufficient for the translation of short, unstructured 50 -UTR mRNAs, defined as ‘‘strong’’ mRNAs, which are essential for normal growth (e.g., b-actin) (De Benedetti and Graff, 2004); once overexpressed, eIF4E is thought to enhance the translation of mRNAs with a long, structured 50 -UTR, known as ‘‘weak’’ mRNAs (Figure 2A). The latter encode proteins such as growth factors (e.g., c-myc and cyclin-D1) and antiapoptotic proteins (e.g., Bcl-2 and survivin) that play the critical roles in malignancy and cell proliferation. Another reported role of eIF4E is to export the mRNAs containing the 4E-sensitivity element (4E-SE) structure, i.e., cyclin-D1, to the cytoplasm (Culjkovic et al., 2007). This process promotes the production of a variety of growth-regulating proteins. Being implicated in various types of cancer, eIF4E has become a major focus of cancer research. The oncogenic activity of eIF4E is related to its phosphorylation by Mnks. In response to extracellular factors, Mnks are activated by Erks and p38 kinases via phosphorylation of two Thr residues present within the activation loop (Thr209 and Thr214 in Mnk1 and Thr244 and Thr249 in Mnk2; vide infra; Jauch et al., 2005, 2006; Waskiewicz et al., 1997). In the Ras/Raf/Erks pathway, stimulated Ras activates Raf, MEK, and Erks in series, with the latter kinases activating the Mnks (Figure 3). The p38 subfamily of MAPKs stimulates Mnk activity via a stress-activated cascade of kinases. Activated Mnks bind to eIF4G, bringing the kinase and substrate into proximity (Pyronnet et al., 1999; Shveygert et al., 2010), thus facilitating the phosphorylation of eIF4E. To date, only two human Mnk genes have been identified, both of which give rise to alternatively spliced isoforms: Mnk1a and Mnk1b, Mnk2a and Mnk2b, respectively (Figure 2B; O’Loghlen et al., 2004; Slentz-Kesler et al., 2000). The four isoforms share similar N-terminal regions that mainly contain a polybasic sequence (PBS) involved in binding to eIF4G, but they differ in their C-termini. The Mnk1a

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 1

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review Figure 1. Roles of eIF4E in Initiating Translation and Exporting mRNAs from the Nucleus to the Cytoplasm (A) Overview of cap-dependent translation initiation: (i) recruitment of the eIF4F complex to the 50 -UTR of the mRNA via an interaction between eIF4E and m7G cap structure, (ii) binding of eIF4B enhances the unwinding of the secondary structures within the 50 -UTR, (iii) recruitment of the 43S preinitiation complex, and (iv) localization of the initiator codon, joining of the large ribosomal unit, and initiation of translation. (B) eIF4E may enhance export to the cytoplasm of certain mRNAs that contain the 4E-SE structure.

eIF4E prevents formation of the eIF4F complex, thus inhibiting cap-dependent translation.

and Mnk2a isoforms have an MAPK-binding site allowing their interaction with Erks and p38 MAPKs (Parra et al., 2005). In contrast to Mnk1a with its high affinity for both kinases, Mnk2a binds predominantly to Erks. The low basal activity of Mnk1a increases after activation by Erks or p38, whereas the constitutive, high basal activity of Mnk2a is affected only slightly by inhibition of the upstream kinases. Notably, Mnk1a possesses a nuclear export sequence (NES) that is related to its presence within the cytoplasm. The b-isoforms are poor substrates for Erks and p38 because they lack the MAPK-binding site within their short C-terminal region. The basal activity of Mnk1b was found to be higher than that of Mnk2b (reviewed in Buxade et al., 2008). Collectively, eIF4E activity is highly regulated by Erks and p38 kinases through Mnk-mediated phosphorylation. In addition, eIF4E is regulated by a family of inhibitory binding proteins, named 4E-BPs, via the phosphatidylinositol 3kinase (PI3K)/Akt/mammalian target of rapamycin complex 1 (mTORC1) pathway (Figure 3; Pause et al., 1994). Extracellular stimulators, i.e., mitogens, cytokines, and growth factors, trigger a cascade of events whereby PI3K activates the 3-phosphoinositide-dependent kinase 1 that promotes the activation of Akt by phosphorylating a residue in its activation loop. Akt inactivates tuberous sclerosis complex 1/2 (TSC1/2) through phosphorylation of TSC2. TSC2 normally acts as the GTPase-activator protein for the GTPase Rheb. Consequently, Rheb-guanosinetriphosphate (Rheb-GTP) accumulates, activating mTORC1 by Rheb-GTP. 4E-BPs are normally bound strongly to eIF4E, but their phosphorylation by mTORC1 releases eIF4E to form the eIF4F complex (Beretta et al., 1996). Since 4E-BPs and eIF4G share the same binding site on eIF4E, the binding of 4E-BPs to

Oncogenic Activity of eIF4E Is Promoted by Its Antiapoptotic Effect Elevated levels of phosphorylated eIF4E are found in human cancer tissues obtained from patients with lung, head, colorectal, and gastric cancers and primary pancreatic ductal adenocarcinoma (Adesso et al., 2013; Fan et al., 2009; Liang et al., 2013). Several experiments on NIH 3T3, Chinese hamster ovary (CHO), and CREF cells showed that high levels of eIF4E were associated with proliferation of cells and an alteration of their morphology (reviewed in Mamane et al., 2004). Wild-type and the Trp56Ala mutant of eIF4E were used to investigate its role in malignancy using a mouse model in which lymphomas were generated from Em-Myc transgenic hematopoietic stem cells (HSCs; Wendel et al., 2007). Mice carrying wild-type eIF4E were found to highly promote Myc-mediated lymphomagenesis, whereas mice reconstituted with the mutant eIF4E were defective in tumor development. Knockdown of eIF4E resulted in a decreased expression of metastasis-related proteins, such as matrix metallopeptidase 9 (MMP9), in HER2-negative MDA-MB-231 and TM15 cells and in a reduced metastatic burden in vivo (Nasr et al., 2013). Taken together, these data provide strong evidence of a role for eIF4E in malignancy. The effects of eIF4E on the cell cycle and apoptosis have also been investigated. Silencing eIF4E led to a prolongation of the G1 phase transition in MDA-MB-231 and TM15 cells and G0/ G1 cell cycle arrest by decreasing the translation of c-myc and MMP9 in nasopharyngeal carcinoma cell lines (Nasr et al., 2013; Wu et al., 2013). A similar effect was observed in UMSCC22B cells, where cell cycle progression was disturbed at a G1-S checkpoint when using eIF4E small interfering RNA (Oridate et al., 2005). In addition, knockdown of eIF4E has been shown to induce apoptotic cell death in MDA-MB-231 cells accompanied by cleavage of the caspase 3 substrate poly (adenosine diphosphate-ribose) polymerase (Yellen et al., 2013). Silencing eIF4E in NIH 3T3 cells significantly reduced the levels of antiapoptotic proteins, including BI-1, dad1, and survivin, which explained the proapoptotic effect of eIF4E depletion (Mamane et al., 2007). Furthermore, it increased the

2 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review Figure 2. Translational Control by Mnks (A) eIF4E and mRNA discrimination. Translation of strong (efficient) mRNAs (blue curve) is independent of eIF4E, whereas translation of ‘‘weak’’ mRNAs (purple curve) requires an increased availability of eIF4E. Strong mRNAs code for vital proteins for normal growth and weak mRNAs for malignancyrelated proteins as in metastasis. (B) Regulatory features of Mnk isoforms. A solid line represents a good affinity for the upstream kinase, whereas a dashed line refers to a poor affinity. Green and blue represent the N terminus and the C terminus, respectively. The following abbreviations were used: KD, kinase domain; MAPK, MAPK binding site. Mnk1b and Mnk2b are poor substrates for the MAPKs since they lack the MAPK binding site. Mnk1a is a good substrate for both while Mnk2a is good substrate only for Erks.

Bax:Bcl-2 ratio in triple-negative breast cancer cells, which induced apoptosis and sensitized the cells to chemotherapy (Silva and Wendel, 2008; Zhou et al., 2011). Notably, an increased sensitivity to fludarabine treatment was achieved when combined with ribavirin, an m7G cap mimetic, in lymphocytes from patients diagnosed with primary chronic lymphocytic leukemia (Martinez-Marignac et al., 2013). It was suggested that ribavirin suppresses the levels of Bcl-2 induced by the single fludarabine treatment, thus mediating cell death. Collectively, the antiapoptotic effects of eIF4E seem important for its oncogenic activity. Role of Mnk Inhibition in Cancer Therapy Mnks phosphorylate eIF4E on Ser209. It has been shown that mice with Ser209Ala eIF4E mutant were resistant to Rasactivated oncogenic transformation (Furic et al., 2010). Mice carrying the ‘‘phosphomimetic’’ Ser209Asp mutant eIF4E showed accelerated tumor onset (Wendel et al., 2007). A constitutively active Mnk1 mutated at Thr332Asp was found to promote tumorigenesis in a similar way to eIF4E. These results correlate the oncogenic activity of eIF4E with its phosphorylation

at Ser209. However, Mnks seem to be dispensable for normal growth (Graff et al., 2007; Ueda et al., 2004). Mnk1/2 double knockout mice have been shown to develop normally without detectable eIF4E phosphorylation (Ueda et al., 2004). More recently, a study suggested that inhibiting eIF4E phosphorylation might differentiate between tumorigenic and normal tissues (Lim et al., 2013). Blast crisis (BC) chronic myeloid leukemia is characterized by the presence of resistant leukemia stem cells (LSCs). This is due to an Akt-mediated elevation of b-catenin signaling in granulocyte macrophage progenitors caused by overexpression and phosphorylation of eIF4E. Since normal HSCs similarly require activated b-catenin signaling, direct inhibition of b-catenin may affect both HSCs and LSCs. However, targeting the Mnk-eIF4E axis was shown to provide an opportunity to selectively prevent BC LSC function in vitro and in vivo without affecting HSC function. Inhibiting Mnk activity, therefore, might present a nontoxic therapeutic opportunity. The roles of Mnks in malignancy have been assessed by several other knock-out/in models. Enhanced expression of Mcl-1, an antiapoptotic protein, was found in Mnk1-expressing lymphomas and correlated with the level of phosphorylated eIF4E (Wendel et al., 2007). Knockdown of Mnk1 using small hairpin RNA decreased the levels of phosphorylated eIF4E and tumor formation in U87MG cells (Ueda et al., 2010). In parallel, attenuated tumor growth was observed in Mnk1/2-doubleknockout PTEN/ mice compared to the parental PTEN/ mice. This finding indicates that inhibiting Mnks might suppress the lymphogenesis driven by the loss of PTEN. Despite an increased understanding of Mnk function and related cancer biology, little progress has been made with validation of the pharmacological target. So far, only a few small molecule inhibitors have been identified. CGP57380 and cercosporamide have served as chemical biological tools, but they lack specificity (Bain et al., 2007; Knauf et al., 2001; Konicek

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 3

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review

Figure 3. eIF4E Is Mainly Regulated by the MAPK Pathways and the PI3K/Akt/mTOR Pathway In response to extracellular stimuli, Mnks (especially Mnk1a) are phosphorylated by Erks or p38 MAPKs. Mnks activate eIF4E by phosphorylating it on Ser209, which appears to promote translation initiation. Furthermore, the availability of eIF4E is regulated by the PI3K/Akt/mTOR pathway. Triggering PI3K activates mTORC1, which in turn phosphorylates 4E-BPs and S6Ks. Hyperphosphorylated 4E-BPs bind weakly to eIF4E, releasing eIF4E to bind eIF4G and thus initiate translation. S6Ks activate eIF4B, which enhances the helicase activity of eIF4A. The Mnk-eIF4E axis may be targeted at different stages: (1) rapalogues inhibit the mTOR pathway, (2) 4EGI-1 impairs the binding of eIF4E to eIF4G, (3) ribavirin reportedly blocks the interaction of eIF4E with the m7G cap structure of mRNAs, and (4) Mnk inhibitors prevent the phosphorylation of eIF4E.

et al., 2011). Besides inhibiting Mnks, CGP57380 inhibits other kinases, including brain-selective serine/threonine kinase 2, casein kinase 1 (CK1; with a similar potency to Mnk1), mitogen-activated protein kinase 1, and proviral integration site 3 (Pim3; Bain et al., 2007). CGP57380 blocked the phosphorylation of eIF4E in HCT-116 and B16 cell lines (Konicek et al., 2011). However, it also inhibited rpS6 phosphorylation in Ba/F3-BcrAbl and K562 cells, thus impairing polysomal assembling and translational initiation (Zhang et al., 2008). CGP57380 induced G1 cell arrest slightly by decreasing the expression of cyclinD1 and cyclin-D3 in these cell lines. Although CGP57380 has been used to reveal the role of eIF4E phosphorylation in cancer cell biology, the information should be interpreted with caution (Buxade et al., 2008). Cercosporamide is another potent Mnk inhibitor (Konicek et al., 2011), but it also targets several other kinases, including Janus kinase 3, glycogen synthase kinase-3b, activin-like kinase-4 and Pim1. It has been shown to inhibit Mnk-mediated eIF4E phosphorylation and induce apoptosis in human cancer cell lines, including HCT-116 and B16 at a concentration above 2.5 mM (Konicek et al., 2011). The compound has also demonstrated antitumor efficacy in animal models. For example, there was significant growth inhibition of human

MV4-11 acute myeloid leukemia (AML) tumors in animals treated twice daily with cercosporamide at 10 mg/kg (Altman et al., 2013). Recent studies have shown that the cytotoxic effects of Mnk inhibitors are cell specific and the potencies seemed to correlate with the levels of phosphorylated eIF4E in cells. For example, MTT experiments on human leukemia U937, MM6, or K562 cells after treatment with cercosporamide showed a dose-dependent suppression of the phosphorylation of eIF4E (Altman et al., 2013). This translated to dose-dependent suppressive effects on leukemic progenitor cell growth. In another study with six breast cancer cell lines, phosphorylated eIF4E was readily detectable in five of them; only MCF-7 cells displayed markedly lower levels (Wheater et al., 2010). Long-term colony-forming assays demonstrated that all the five cell lines having high levels of phosphorylated eIF4E were highly sensitive to CGP57380. These included two ERa-positive lines as well as a HER2/ERa/ PR-negative line that carries an activating k-RAS mutation. In contrast, no inhibition of proliferation was detected in MCF-7 cells. The antiproliferative effects of CGP57380 in breast cancer cells seemed primarily cytostatic, rather than cytotoxic, potentially due to the downregulation of cyclin-D1 synthesis (Wheater et al., 2010).

4 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review Combination of Targeting Mnk-eIF4E and Other Pathways Mnk inhibitors have shown enhanced antitumor activity when combined with other protein kinase inhibitors or chemotherapeutic agents. Although CGP57380 alone caused a partial growth inhibition and a limited proapoptotic effect on cutaneous T cell lymphoma cells, its combination with rapamycin, an mTOR inhibitor, achieved complete cell growth inhibition and induced apoptotic cell death (Marzec et al., 2011). The combined treatment also inhibited cell cycle progression and suppressed the growth of prostate cancer cells (Bianchini et al., 2008). The combination of CGP57380 with imatinib, a Bcr-Abl inhibitor, improved its effect on cell-cycle arrest and caspase-3-mediated cell death in Ba/F3-Bcr-Abl and K562 cells (Zhang et al., 2008). Similarly, treatment of pancreatic cancer cell lines MiaPaCa2 and PT45P1 with CGP57380 in combination with gemcitabine caused a greater apoptotic cell death when compared to the use of either CGP57380 or gemcitabine alone (Adesso et al., 2013), again suggesting the synergistic effects of these compounds. A similar picture has emerged from clinical trials of therapeutic drugs targeting the Mnk-eIF4E axis. eIF4E-specific antisense oligonucleotides (4E-ASOs) caused a decrease in the expression of proteins important for malignancy (e.g., cyclin-D1, c-myc, and Bcl-2), induced apoptosis, and prevented tumor growth (Graff et al., 2007). However, LY2275796, a 4E-ASO currently in phase I/II clinical trials, did not give rise to any tumor response even with a reduction in eIF4E mRNA of 80% within posttreatment tumor biopsies (Hong et al., 2011). The results may reflect a less robust downregulation of eIF4E in humans resulting in a cytostatic effect. Based on this, the use of LY2275796 as a chemo-sensitizer with chemotherapeutic agents has been proposed. Likewise, ribavirin inhibited the activity of eIF4E and slowed tumor growth in a xenograft mouse model of eIF4E-dependent human squamous cell carcinoma (Kentsis et al., 2004). It inhibited proliferation of multiple myeloma cells by 43% and reduced the expression of cyclin-D1 and c-myc (Li et al., 2013). Nevertheless, its efficacy in the M4 and M5 subtypes of AML patients was counteracted by a molecular resistance, despite the attenuation of eIF4E levels (Assouline et al., 2009; Borden, 2011). The utility of combining ribavirin with a cytotoxic chemotherapeutic agent is currently being explored in clinical trials of poor prognosis AML. Being implicated in the deregulation of protein synthesis, eIF4E exhibits an oncogenic profile associated with its phosphorylation by Mnks. The role of Mnks in tumorigenesis has been confirmed by several cancer models. Mnks have emerged as potential effective and nontoxic anticancer targets. To date, however, little progress has been made in the pharmacologic target validation. Identification of highly potent and selective Mnk inhibitors would be of importance to further our understanding of Mnk-related tumor biology and therapeutic applications. Structural Features of Mnks Mnk kinase domain exists in the bilobal scaffold common to protein kinases: the ATP-binding site is located in a cleft between the N-lobe and the C-lobe (Figure 4A; Jauch et al., 2005). The N-lobe is a small hydrophobic lobe formed by antiparallel b sheets and a regulatory aC-helix (Jauch et al., 2005, 2006).

Highly conserved sequence motifs are buried within the N-lobe, including a Gly-rich loop and an AxK motif in the b3 strand. While there is 78% sequence similarity between Mnk1 and Mnk2 in their catalytic domains, the ATP binding site of Mnk2 is more accessible to ATP or other substrates due to the rotation of the N-lobe of Mnk2 by a further 10 . The N-lobe is connected to the C-lobe by a segment called the hinge region, which begins at the gatekeeper residue Phe159 (Mnk2 numbering is used throughout the review unless otherwise stated). The C-lobe commonly encompasses six helices and comprises the catalytic loop and the activation segment. The activation segment consists of the magnesium binding loop, the activation loop, and the P+1 loop (Jauch et al., 2005, 2006). The magnesium binding loop begins, uniquely, with a DFD motif, instead of the DFG motif common to all other protein kinases. The activation loop also harbors the phosphorylation sites in Mnks, Thr244, and Thr249. Furthermore, three Mnk-specific insertions are present within the catalytic domain (Figures 4A and 4B): insertion I1 within the C terminus of the DFD motif, insertion I2 within the aEF/aF loop, and insertion I3 in the N terminus of the aG-helix. Mnks Favor an Autoinhibitory ‘‘DFD-Out’’ Conformation One of the mechanisms that regulate kinase catalytic activity is the conformational change of the DFG/D motif. In the inactive state of Mnk2, the DFD motif adopts an ‘‘out’’ conformation, in which DFD-Phe227 (Phe192 in Mnk1) inserts into the ATP-binding site, thus preventing ATP from entering this binding pocket (Figure 4C). To date, the only wild-type Mnk1 and Mnk2 crystal structures observed are in the DFD-out conformation. This observation of DFG/D-out conformations is rare given that only about 3% of kinases were captured in that conformation (Kufareva and Abagyan, 2008; Zuccotto et al., 2010). For the activation of Mnks, the DFD motif is expected to flip to an ‘‘in’’ conformation where the DFD-Phe227 moves away from the ATP binding site. The active and the inactive states of kinases might also be described in terms of spines, i.e., a regulatory spine (R-spine) and a catalytic spine (C-spine; Kornev et al., 2006; Kornev and Taylor, 2010; Taylor and Kornev, 2011). The spines can be observed utilizing a local spatial-patterns alignment, a bioinformatics method for comparing the spatial arrangements of protein structural elements without considering the sequence or the geometry of the main chains of the protein. Both spines exist only in the active conformation and are disassembled in the inactive conformation. The R-spine bridges the N-lobe and the C-lobe in the active state and is triggered by the phosphorylation of the activation loop. The C-spine also links both N- and C-lobes and is only completed by the adenine ring of the ATP. Similar to other kinases, Mnks are expected to form the R-spine by forming hydrophobic interactions between four residues, two from the N-lobe (aC-helix-Leu133 and b4 strand-Leu145) and two from the C-lobe (HRD-His203 and DFD-Phe227). The C-spine will include Ala111 from the AxK motif in the b3 strand, which makes direct contact with the adenosyl ring of ATP. The Mnk DFD-out conformation is stabilized by various interactions. In Mnk2, a hydrogen bond links the amino of the b3-Lys113 to the carbonyl of the DFD-Phe227, stabilizing the DFD-out conformation (Jauch et al., 2005). The interactions

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 5

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review

Figure 4. Structure of Mnks (A) Stereoview of the structure of Mnk2 in the DFD-out conformation based on a homology model that uses the X-ray crystallographic structure (PDB ID 2AC3) as the template (Hou et al., 2013). The hinge region backbone is shown in yellow, the three insertions in red. The DFD motif and the Phe gatekeeper are in capped sticks. (legend continued on next page)

6 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review between HRD-Arg204 (Arg169 in Mnk1) and the aF-helixAsp273 (Asp238 in Mnk1) also help in favoring the inactive Mnk conformations. It is believed that Mnk1 DFD-out conformation is more stable compared to Mnk2 due to several additional interactions. Among these, the relocation of I2-Phe230 to the DFG/D-in pocket (where the DFD-Phe192 is situated in the DFD-in conformation) is particularly interesting. However, such movement is unlikely in Mnk2 as I2 is located farther from the aforementioned hydrophobic pocket. The role of the unique DFD-Asp228 in stabilizing the DFD-out conformation has also been studied. Mutating the DFD-Asp228 with Gly in the Mnk2 kinase region showed that Mnk2 can now adopt both DFG-in and DFG-out conformations (Jauch et al., 2005), presumably due to the flexibility gained by the presence of Gly. This was further confirmed by molecular dynamics (MD) simulation studies, where the Mnk-DFG model readily sampled both DFG-in and DFG-out conformations (Hou et al., 2013). It has been shown that lacking a side chain, Gly demonstrates higher flexibility as a bipositional switch that positions the DFG motif correctly for the active or inactive state (Kornev et al., 2006). DFD-Asp228 is not capable of achieving such flexibility and this may manifest as a crucial factor in favoring the DFD-out conformation for Mnks. Observation of Mnk1 DFD-Asp193 hydrogen bonding with other residues in the DFD-out conformation, stabilizing the inactive state, only supports this notion. A comparative study where the DFG-Gly is mutated to DFD-Asp in a well-characterized kinase, e.g., c-Abl, might provide further insights into the exact role of the DFD-Asp in the DFD-in/-out switch. Mnk-specific insertions also contribute to the stabilization of the inactive conformation. Residues of insertion I2, e.g., Glu228 and Phe230, are thought to be involved in repositioning the activation segment into the DFD-out conformation for Mnk1 (Jauch et al., 2006). MD simulations have been used to investigate the role of insertion I1 (which is only four residues away from the DFD motif) in Mnk2 during the conformational changes (Hou et al., 2013). The authors demonstrated that an interaction involving DFD-Asp226 and I1-Lys234 was stabilizing the DFD-out conformation, highlighting the importance of I1 for a favorable autoinhibitory state. Activation of Mnks As with other kinases, Mnks are expected to become more organized upon activation via phosphorylation, in preparation for catalysis. It has been suggested that the interactions between the ATP phosphates and the HRD-Arg204, which in turn disrupt HRD-Arg204’s interaction with the aF-helix-Asp273 (Figure 4C), are important for triggering the switch from the inactive to active conformation (Jauch et al., 2006; Taylor and Kornev, 2011). Release of the aF-helix-Asp273 may then destabilize the extruded conformation of the activation segment, which in turn flips inward, facilitating a restoration of the interaction between subdomain VIII and the aF/aG/C-loop region (Jauch et al.,

2005). Additionally, the positioning of the aC-helix has also been shown to be a key regulatory element in protein kinase activation process (Taylor and Kornev, 2011) and a similar behavior might be expected from Mnks. Further investigations into the dynamical events of the activation process of Mnks would reveal any intermediate conformations that can be exploited for designing selective inhibitors. As such, we are currently investigating these dynamic events using computational simulations. ATP-Competitive Kinase Inhibitors The strategies being followed for the discovery of new anticancer agents have shifted from randomly high-throughput screening to molecularly targeted approaches. Imatinib, the first kinase inhibitor to reach the market as an anticancer drug in 2001, marked the beginning of a new era for cancer treatment—targeted cancer therapy. To date, fourteen kinase inhibitors have received regulatory approval, making a considerable impact on cancer therapy (Liu et al., 2013). The majority of kinase inhibitors are reversible ATP-competitors and are termed type I and type II inhibitors. Typically, type I inhibitors mimic the interactions of the adenine ring of ATP with the hinge region residues of their respective kinases (Liu and Gray, 2006). They normally bind to an active kinase structure that is capable of binding ATP. The drawback of this type of inhibitor is that they tend to inhibit other kinases due to the highly conserved structure of ATP-binding domain. This might result in off-target side effects and toxicity. A few type I inhibitors such as VX-745 and CI-1033 have been designed to overcome this problem by interacting with less conserved residues within the binding site (Karaman et al., 2008). Type II inhibitors, e.g., imatinib and BIRB-796, bind to the inactive DFG-out structure. They bind to the ATP binding site and to an adjacent hydrophobic binding pocket (vide infra) that is made accessible by the rearrangement of the DFG/D motif to the DFG/D-out conformation. As there are substantial variations among the inactive conformations of kinases compared to their ATP-activated structures, type II inhibitors are expected to have higher selectivity compared to type I inhibitors (Zhang et al., 2009). The DFD-out inactive form of Mnks can also be readily exploited for designing highly selective inhibitors. Covalent Irreversible Inhibitors Development of covalent agents has been treated cautiously due to their indiscriminate and irreversible binding to proteins other than the target, therefore generating toxicity. More recently, there has been renewed interest in the development of irreversible kinase inhibitors that form covalent bond with nucleophilic residue in the ATP-binding domain. These inhibitors contain Michael acceptor-type substituent at an appropriate position on a chemical scaffold and exploit the presence of a cysteine residue in the ATP-binding site to establish the addition product when bound to the enzyme (Singh et al., 2011). By

(B) Sequence alignment of kinase domains of Mnks, FLT3, and c-KIT. The DFG/D motif is shown in red, the phosphorylation sites in green, and the three insertions in line-edged boxes. The light blue background highlights conserved residues among all four kinases, and light orange the residues conserved only between Mnk1/2. (C) Stereoview of the superposition of the DFD-in (light blue) and -out (green) conformations of Mnk2. The homology model is built to include the missing activation loop and insertion 3 (I3) in the X-ray crystal structure (PDB ID 2HW7). Phe227-in represents Phe227 in the DFD-in conformation of Mnk2. The hinge region is shown in yellow. All figures are generated using Pymol v.1.5.0.4.

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 7

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review

Figure 5. Known and Proposed Mnk Inhibitors (A) Stereoview of X-ray structure of the binding mode of staurosporine to Mnk2 in the DFG-in conformation (PDB ID 2HW7). Mnk2 is shown as light-brown ribbons. Staurosporine is shown in capped sticks. Hydrogen bonds are shown as dashed lines. (B) The chemical structures of known Mnk inhibitors and proposed Mnk2 inhibitors (PI-1 and PI-2).

modulation and fine tuning of chemical functionalities, covalent irreversible inhibitors can be developed as effective therapeutic agents with the limited off-target side effects (Serafimova et al., 2012). Irreversible inhibitors have a number of potential advantages, including prolonged pharmacodynamics, high potency, and an ability to overcome the problem of drug resistance when compared to their reversible counterparts. Design of Mnk Inhibitors Despite the increased understanding of the role of Mnks in cancer, little progress has been made in the rational design of inhibitors. Oyarzabal et al. (2010) employed a fragment-oriented virtual screening protocol to identify several compounds with potent inhibitory activity against Mnks. One of these has been shown to demonstrate a selective profile against 22 other protein kinases. Another analog-design approach was used to develop covalent Mnk inhibitors, although their specificity was not assessed (Xu et al., 2013). The lack of progress in Mnk-targeted drug discovery can be attributed to the absence of thorough understanding of structural details. No structural information of the

activated Mnks is yet available. A staurosporine-bound crystal structure of Mnk Asp228Gly mutant showed that the polycyclic ring of staurosporine occupies the ATP cleft where two hydrogen bonds are formed between the 1-NH and 5-O atoms of staurosporine and Glu160 and Met162, respectively, in the hinge region (Figure 5A; Jauch et al., 2006). These hinge region interactions demonstrate a known binding mode exploited by other kinase inhibitors. However, due to the Asp228Gly mutation, any conformational changes that Asp228 might induce can remain hidden. Thus, we generated a homology model of the wild-type DFD-in conformation of Mnk2 (Figure 6) using Modeler (Eswar et al., 2006). This model was selected based on the Modeler score and Ramachandran plots, from an ensemble of 1,000 initial models. The ATP-binding site of the chosen DFD-in model was mapped to reveal electrostatic and steric features as shown in Figure 6A. The map details vital differences between the features of the DFD-in and -out conformations that can be exploited to tailor selective inhibitor designs. For instance, the map clearly illustrates the size and shape of the hydrophobic pocket that is generated in the DFD-out conformation (Figure 6B), when

8 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review

Figure 6. Hydrogen Bond Donor/Acceptor and Hydrophobic Area Mapping of the Binding Site of Mnk2 The Mnk2 DFD-in conformation is shown in (A), and the DFD-out homology model in (B). Phe227 of the DFD motif is shown in black and the protein backbone is in ribbons. The yellow solid surface depicts hydrophobic areas where hydrophobic moieties of ligands could be positioned. The red and blue mesh indicates areas where hydrogen bond acceptor and donor functionalities can be placed. The figures were generated using Site Map and Maestro (Schro¨dinger Suite 2012).

Phe227 moves out of the DFD-in position. Type II inhibitors may be designed to take advantage of this hydrophobic pocket and the unusually stable DFD-out conformation. Similar to hydrophobic areas, the maps also illustrate the hydrogen bond donor/ acceptor areas in the vicinity, including those of the hinge region. To examine the binding modes of known Mnk inhibitors, the homology model of the DFD-in conformation was used to dock and score CGP57380, using the induced fit docking (IFD) protocol (Sherman et al., 2006). The docking pose reveals the formation of strong hydrogen bonds between the 1-NH and 2-N groups of the pyrazolo[3,4-d]pyrimidine system of the CGP57380 and the backbone of Glu160 and Met162 residues in the hinge region of Mnk2 (Figure 7A). As demonstrated previously, the model attests to the necessity of the 1-NH moiety and its removal by methylation, e.g., compound SHN-093 (Figure 5B), would abolish the inhibition (Buxade´ et al., 2005). However, we note some dissimilarity between the current binding mode and our previous one at the site of hinge interactions (Hou et al., 2012). Previously, the hinge interactions were shown through 1-NH, 2-N, and 3-NH of pyrazolo[3,4-d]pyrimidine. The difference is most likely due to the introduction of flexibility to the protein structure by using the IFD protocol, which allows the identification of optimal binding modes. Additionally, in the current binding mode, we note that the p-fluorophenyl moiety is engaged in a p-p interaction with the gatekeeper Phe159, providing further stabilization. Improving Selectivity for Mnks Information on electrostatic and steric features of the ATP binding site of Mnks can be used together with those of other kinases to direct selective inhibitor design. Superimposing the binding site map with the docking pose of CGP57380 suggests that a (piperazin-1-yl)acetamide moiety can be introduced at the C4position, while replacing the C3-aniline moiety by a methyl group (PI-1; Figure 5B). The binding mode generated by IFD demon-

strates that PI-1 interacts with Ser231 through the amino group of the piperazine moiety, while maintaining the hinge interactions seen with CGP57380 (Figure 7B). Being less conserved among kinases, Ser231 in the activation loop of Mnk2 provides a convenient stepping stone to increase potency and selectivity for Mnk2. Although the IFD protocol was used, significant movements of the protein backbone at the ATP-binding site were not observed. This minimum reorganization required by the protein compared to the binding conformation of a verified Mnk inhibitor, CGP57380, provides additional confidence. Compound PI-1 may be modified to increase selectivity even further by targeting another partially conserved residue, Ser166, which is positioned adjacent to the hinge region. The proposed compound PI-2 has been designed to target Ser166 by introducing an additional butan-2-one moiety at the 4C-NH of pyrazolo[3,4-d]pyrimidin (Figure 5B). Our binding mode suggests that while PI-2 interacts with the Ser231 in a similar manner to PI-1, the butan-2-one moiety seems engaged in a bifurcated hydrogen bond with the hydroxyl and the backbone amino groups of the Ser166 (Figure 7C). As PI-2 targets two residues that are partially conserved among kinases, we expect it to be more selective than PI-1. The fully conserved DFD-Asp228 was not targeted in our designs, as Asp228 in the current energy-minimized homology model of the DFD-in state resolved to an inaccessible conformation. However, as the activation loops of kinases are highly dynamic, we cannot exclude the possibility of a DFD-in conformation where Asp228 would be accessible to interaction with inhibitors. We are currently investigating this possibility using more extensive computer simulations. Specific protein residues may also be targeted for interaction by irreversible inhibitors to make a covalent complex with the protein. In the case of Mnks, this can be achieved by targeting cysteine(s) of the binding pocket, specifically Cys190/Cys225 of Mnk1/Mnk2. For instance, resorcylic acid analogs were recently identified as covalent Mnk inhibitors targeting

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 9

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review

Figure 7. Binding Modes of CGP57380, PI-1, and PI-2 to Mnk2 (A) Binding mode of CGP57380 to ATP-binding site and the schematic of protein-inhibitor interactions. (B) Binding mode of the designed inhibitor PI-1. PI-1 interacts with Ser231, a partially conserved residue among kinases. (C) Binding mode of PI-2. PI-2 interacts with Ser166 and 231, both partially conserved residues. Mnk2 is shown in light blue ribbons with the bound inhibitors in capped sticks. Dashed lines are hydrogen bonds. The binding modes were generated computationally using the IFD protocol (Schro¨dinger Suite 2012). A homology model of the DFD-in conformation based on an X-ray structure (PDB ID 2HW7) provided the protein atom coordinates. The ligand-interaction-diagram tool (Schro¨dinger Suite 2012) was used to produce the interaction schematics.

Cys225, one of the readily accessible cysteine residues for covalent interaction in Mnk2 (Xu et al., 2013). The design approach was based on the binding mode of hypothemycin with Erk2 (Protein Data Bank [PDB] ID 3C9W). Alternatively, a fragment-based approach could be employed for designing selective Mnk inhibitors. This can be achieved by incorporating an electrophilic group to the scaffold of an inhibitor showing a submicromolar affinity for Mnks, e.g., CGP57380. As Cys190/ Cys225 of Mnk1/Mnk2 is conserved in 46 other kinases, hybrid

design approaches could be employed to reduce off-target effects (Cohen et al., 2005). For instance, forming p-p interaction with the gatekeeper and hydrogen bonding with unique or partially conserved residues might offer room for improving selectivity. Conclusions Significant progress has been made in understanding of the involvement of Mnk-eIF4E in oncogenic transformation and

10 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review development. Strong biological evidence supports Mnks as anticancer targets, but their pharmacological validation is still in its infancy. Comprehension of Mnk kinase structural elements would rationally guide inhibitor design and selectivity optimization. Given the complexity of cancer, a combination of Mnk inhibitor with other critical pathways targeting agent(s) might be considered as an additional therapeutic approach. However, a promiscuous multitarget activity might pose considerable risks due to unforeseeable side effects and toxicity. A rationally designed combination therapy based on the selective inhibition on a specific set of kinase targets would reduce these risks. For these reasons, the discovery of highly selective Mnk inhibitors will be a major step toward drug development as singletarget agents as well as combination therapeutics for cancer. ACKNOWLEDGMENTS This study is supported by Australia Government National Health and Medical Research Council (research grant 1050825).

Eswar, N., Webb, B., Marti-Renom, M.A., Madhusudhan, M., Eramian, D., Shen, M.Y., Pieper, U., and Sali, A. (2006). Comparative protein structure modeling using Modeller. Curr. Protoc. Bioinformatics 5, 5.6. Fan, S., Ramalingam, S.S., Kauh, J., Xu, Z., Khuri, F.R., and Sun, S.Y. (2009). Phosphorylated eukaryotic translation initiation factor 4 (eIF4E) is elevated in human cancer tissues. Cancer Biol. Ther. 8, 1463–1469. Feoktistova, K., Tuvshintogs, E., Do, A., and Fraser, C.S. (2013). Human eIF4E promotes mRNA restructuring by stimulating eIF4A helicase activity. Proc. Natl. Acad. Sci. USA 110, 13339–13344. Furic, L., Rong, L., Larsson, O., Koumakpayi, I.H., Yoshida, K., Brueschke, A., Petroulakis, E., Robichaud, N., Pollak, M., Gaboury, L.A., et al. (2010). eIF4E phosphorylation promotes tumorigenesis and is associated with prostate cancer progression. Proc. Natl. Acad. Sci. USA 107, 14134–14139. Graff, J.R., Konicek, B.W., Vincent, T.M., Lynch, R.L., Monteith, D., Weir, S.N., Schwier, P., Capen, A., Goode, R.L., Dowless, M.S., et al. (2007). Therapeutic suppression of translation initiation factor eIF4E expression reduces tumor growth without toxicity. J. Clin. Invest. 117, 2638–2648. Hong, D.S., Kurzrock, R., Oh, Y., Wheler, J., Naing, A., Brail, L., Callies, S., Andre´, V., Kadam, S.K., Nasir, A., et al. (2011). A phase 1 dose escalation, pharmacokinetic, and pharmacodynamic evaluation of eIF-4E antisense oligonucleotide LY2275796 in patients with advanced cancer. Clin. Cancer Res. 17, 6582–6591.

REFERENCES

Hou, J., Lam, F., Proud, C., and Wang, S. (2012). Targeting Mnks for cancer therapy. Oncotarget 3, 118–131.

Adesso, L., Calabretta, S., Barbagallo, F., Capurso, G., Pilozzi, E., Geremia, R., Delle Fave, G., and Sette, C. (2013). Gemcitabine triggers a pro-survival response in pancreatic cancer cells through activation of the MNK2/eIF4E pathway. Oncogene 32, 2848–2857.

Hou, J., Teo, T., Sykes, M.J., and Wang, S. (2013). Insights into the importance of DFD-motif and insertion I1 in stabilizing the DFD-out conformation of Mnk2 kinase. ACS Med. Chem. Lett. 4, 736–741.

Altman, J.K., Szilard, A., Konicek, B.W., Iversen, P.W., Kroczynska, B., Glaser, H., Sassano, A., Vakana, E., Graff, J.R., and Platanias, L.C. (2013). Inhibition of Mnk kinase activity by cercosporamide and suppressive effects on acute myeloid leukemia precursors. Blood 121, 3675–3681. Assouline, S., Culjkovic, B., Cocolakis, E., Rousseau, C., Beslu, N., Amri, A., Caplan, S., Leber, B., Roy, D.-C., Miller, W.H., Jr., and Borden, K.L. (2009). Molecular targeting of the oncogene eIF4E in acute myeloid leukemia (AML): a proof-of-principle clinical trial with ribavirin. Blood 114, 257–260. Bain, J., Plater, L., Elliott, M., Shpiro, N., Hastie, C.J., McLauchlan, H., Klevernic, I., Arthur, J.S.C., Alessi, D.R., and Cohen, P. (2007). The selectivity of protein kinase inhibitors: a further update. Biochem. J. 408, 297–315. Beretta, L., Gingras, A.C., Svitkin, Y.V., Hall, M.N., and Sonenberg, N. (1996). Rapamycin blocks the phosphorylation of 4E-BP1 and inhibits cap-dependent initiation of translation. EMBO J. 15, 658–664. Bianchini, A., Loiarro, M., Bielli, P., Busa`, R., Paronetto, M.P., Loreni, F., Geremia, R., and Sette, C. (2008). Phosphorylation of eIF4E by MNKs supports protein synthesis, cell cycle progression and proliferation in prostate cancer cells. Carcinogenesis 29, 2279–2288. Borden, K.L.B. (2011). Targeting the oncogene eIF4E in cancer: from the bench to clinical trials. Clin. Invest. Med. 34, E315–E319. Buxade´, M., Parra, J.L., Rousseau, S., Shpiro, N., Marquez, R., Morrice, N., Bain, J., Espel, E., and Proud, C.G. (2005). The Mnks are novel components in the control of TNF a biosynthesis and phosphorylate and regulate hnRNP A1. Immunity 23, 177–189. Buxade, M., Parra-Palau, J.L., and Proud, C.G. (2008). The Mnks: MAP kinaseinteracting kinases (MAP kinase signal-integrating kinases). Front. Biosci. 13, 5359–5373. Cohen, M.S., Zhang, C., Shokat, K.M., and Taunton, J. (2005). Structural bioinformatics-based design of selective, irreversible kinase inhibitors. Science 308, 1318–1321.

Jauch, R., Ja¨kel, S., Netter, C., Schreiter, K., Aicher, B., Ja¨ckle, H., and Wahl, M.C. (2005). Crystal structures of the Mnk2 kinase domain reveal an inhibitory conformation and a zinc binding site. Structure 13, 1559–1568. Jauch, R., Cho, M.K., Ja¨kel, S., Netter, C., Schreiter, K., Aicher, B., Zweckstetter, M., Ja¨ckle, H., and Wahl, M.C. (2006). Mitogen-activated protein kinases interacting kinases are autoinhibited by a reprogrammed activation segment. EMBO J. 25, 4020–4032. Karaman, M.W., Herrgard, S., Treiber, D.K., Gallant, P., Atteridge, C.E., Campbell, B.T., Chan, K.W., Ciceri, P., Davis, M.I., Edeen, P.T., et al. (2008). A quantitative analysis of kinase inhibitor selectivity. Nat. Biotechnol. 26, 127–132. Kentsis, A., Topisirovic, I., Culjkovic, B., Shao, L., and Borden, K.L. (2004). Ribavirin suppresses eIF4E-mediated oncogenic transformation by physical mimicry of the 7-methyl guanosine mRNA cap. Proc. Natl. Acad. Sci. USA 101, 18105–18110. Knauf, U., Tschopp, C., and Gram, H. (2001). Negative regulation of protein translation by mitogen-activated protein kinase-interacting kinases 1 and 2. Mol. Cell. Biol. 21, 5500–5511. Konicek, B.W., Stephens, J.R., McNulty, A.M., Robichaud, N., Peery, R.B., Dumstorf, C.A., Dowless, M.S., Iversen, P.W., Parsons, S., Ellis, K.E., et al. (2011). Therapeutic inhibition of MAP kinase interacting kinase blocks eukaryotic initiation factor 4E phosphorylation and suppresses outgrowth of experimental lung metastases. Cancer Res. 71, 1849–1857. Kornev, A.P., and Taylor, S.S. (2010). Defining the conserved internal architecture of a protein kinase. Biochim. Biophys. Acta 1804, 440–444. Kornev, A.P., Haste, N.M., Taylor, S.S., and Eyck, L.F. (2006). Surface comparison of active and inactive protein kinases identifies a conserved activation mechanism. Proc. Natl. Acad. Sci. USA 103, 17783–17788. Kufareva, I., and Abagyan, R. (2008). Type-II kinase inhibitor docking, screening, and profiling using modified structures of active kinase states. J. Med. Chem. 51, 7921–7932.

Culjkovic, B., Topisirovic, I., and Borden, K.L.B. (2007). Controlling gene expression through RNA regulons: the role of the eukaryotic translation initiation factor eIF4E. Cell Cycle 6, 65–69.

Li, S., Fu, J., and Suzanne Lentzsch, M. (2013). Cap-dependent protein translation initiation in multiple myeloma: an attractive target for therapy. In Genetic and Molecular Epidemiology of Multiple Myeloma, S. Lentzsch, ed. (New York: Springer), pp. 43–57.

De Benedetti, A., and Graff, J.R. (2004). eIF-4E expression and its role in malignancies and metastases. Oncogene 23, 3189–3199.

Liang, S., Guo, R., Zhang, Z., Liu, D., Xu, H., Xu, Z., Wang, X., and Yang, L. (2013). Upregulation of the eIF4E signaling pathway contributes to the

Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved 11

Please cite this article in press as: Diab et al., MAP Kinase-Interacting Kinases—Emerging Targets against Cancer, Chemistry & Biology (2014), http:// dx.doi.org/10.1016/j.chembiol.2014.01.011

Chemistry & Biology

Review progression of gastric cancer, and targeting eIF4E by perifosine inhibits cell growth. Oncol. Rep. 29, 2422–2430. Lim, S., Saw, T.Y., Zhang, M., Janes, M.R., Nacro, K., Hill, J., Lim, A.Q., Chang, C.-T., Fruman, D.A., Rizzieri, D.A., et al. (2013). Targeting of the MNK-eIF4E axis in blast crisis chronic myeloid leukemia inhibits leukemia stem cell function. Proc. Natl. Acad. Sci. USA 110, E2298–E2307. Liu, Y., and Gray, N.S. (2006). Rational design of inhibitors that bind to inactive kinase conformations. Nat. Chem. Biol. 2, 358–364. Liu, Q., Sabnis, Y., Zhao, Z., Zhang, T., Buhrlage, S.J., Jones, L.H., and Gray, N.S. (2013). Developing irreversible inhibitors of the protein kinase cysteinome. Chem. Biol. 20, 146–159. Mamane, Y., Petroulakis, E., Rong, L., Yoshida, K., Ler, L.W., and Sonenberg, N. (2004). eIF4E—from translation to transformation. Oncogene 23, 3172– 3179. Mamane, Y., Petroulakis, E., Martineau, Y., Sato, T.-A., Larsson, O., Rajasekhar, V.K., and Sonenberg, N. (2007). Epigenetic activation of a subset of mRNAs by eIF4E explains its effects on cell proliferation. PLoS ONE 2, e242. Martinez-Marignac, V., Shawi, M., Pinedo-Carpio, E., Wang, X., Panasci, L., Miller, W., Pettersson, F., and Aloyz, R. (2013). Pharmacological targeting of eIF4E in primary CLL lymphocytes. Blood Cancer J. 3, e146. Marzec, M., Liu, X., Wysocka, M., Rook, A.H., Odum, N., and Wasik, M.A. (2011). Simultaneous inhibition of mTOR-containing complex 1 (mTORC1) and MNK induces apoptosis of cutaneous T-cell lymphoma (CTCL) cells. PLoS ONE 6, e24849. Nasr, Z., Robert, F., Porco, J.A., Jr., Muller, W.J., and Pelletier, J. (2013). eIF4F suppression in breast cancer affects maintenance and progression. Oncogene 32, 861–871. O’Loghlen, A., Gonza´lez, V.M., Pin˜eiro, D., Pe´rez-Morgado, M.I., Salinas, M., and Martı´n, M.E. (2004). Identification and molecular characterization of Mnk1b, a splice variant of human MAP kinase-interacting kinase Mnk1. Exp. Cell Res. 299, 343–355. Oridate, N., Kim, H.-J., Xu, X., and Lotan, R. (2005). Growth inhibition of head and neck squamous carcinoma cells by small interfering RNAs targeting eIF4E or cyclin D1 alone or combined with cisplatin. Cancer Biol. Ther. 4, 318–323. Oyarzabal, J., Zarich, N., Albarran, M.I., Palacios, I., Urbano-Cuadrado, M., Mateos, G., Reymundo, I., Rabal, O., Salgado, A., Corrionero, A., et al. (2010). Discovery of mitogen-activated protein kinase-interacting kinase 1 inhibitors by a comprehensive fragment-oriented virtual screening approach. J. Med. Chem. 53, 6618–6628. Parra, J.L., Buxade´, M., and Proud, C.G. (2005). Features of the catalytic domains and C termini of the MAPK signal-integrating kinases Mnk1 and Mnk2 determine their differing activities and regulatory properties. J. Biol. Chem. 280, 37623–37633. Pause, A., Belsham, G.J., Gingras, A.-C., Donze´, O., Lin, T.-A., Lawrence, J.C., Jr., and Sonenberg, N. (1994). Insulin-dependent stimulation of protein synthesis by phosphorylation of a regulator of 50 -cap function. Nature 371, 762–767. Pyronnet, S., Imataka, H., Gingras, A.C., Fukunaga, R., Hunter, T., and Sonenberg, N. (1999). Human eukaryotic translation initiation factor 4G (eIF4G) recruits mnk1 to phosphorylate eIF4E. EMBO J. 18, 270–279. Rhoads, R.E. (1988). Cap recognition and the entry of mRNA into the protein synthesis initiation cycle. Trends Biochem. Sci. 13, 52–56.

Shveygert, M., Kaiser, C., Bradrick, S.S., and Gromeier, M. (2010). Regulation of eukaryotic initiation factor 4E (eIF4E) phosphorylation by mitogen-activated protein kinase occurs through modulation of Mnk1-eIF4G interaction. Mol. Cell. Biol. 30, 5160–5167. Silva, R.L.A., and Wendel, H.G. (2008). MNK, EIF4E and targeting translation for therapy. Cell Cycle 7, 553–555. Singh, J., Petter, R.C., Baillie, T.A., and Whitty, A. (2011). The resurgence of covalent drugs. Nat. Rev. Drug Discov. 10, 307–317. Slentz-Kesler, K., Moore, J.T., Lombard, M., Zhang, J., Hollingsworth, R., and Weiner, M.P. (2000). Identification of the human Mnk2 gene (MKNK2) through protein interaction with estrogen receptor b. Genomics 69, 63–71. Taylor, S.S., and Kornev, A.P. (2011). Protein kinases: evolution of dynamic regulatory proteins. Trends Biochem. Sci. 36, 65–77. Ueda, T., Watanabe-Fukunaga, R., Fukuyama, H., Nagata, S., and Fukunaga, R. (2004). Mnk2 and Mnk1 are essential for constitutive and inducible phosphorylation of eukaryotic initiation factor 4E but not for cell growth or development. Mol. Cell. Biol. 24, 6539–6549. Ueda, T., Sasaki, M., Elia, A.J., Chio, I.I.C., Hamada, K., Fukunaga, R., and Mak, T.W. (2010). Combined deficiency for MAP kinase-interacting kinase 1 and 2 (Mnk1 and Mnk2) delays tumor development. Proc. Natl. Acad. Sci. USA 107, 13984–13990. Waskiewicz, A.J., Flynn, A., Proud, C.G., and Cooper, J.A. (1997). Mitogenactivated protein kinases activate the serine/threonine kinases Mnk1 and Mnk2. EMBO J. 16, 1909–1920. Wendel, H.G., Silva, R.L.A., Malina, A., Mills, J.R., Zhu, H., Ueda, T., Watanabe-Fukunaga, R., Fukunaga, R., Teruya-Feldstein, J., Pelletier, J., and Lowe, S.W. (2007). Dissecting eIF4E action in tumorigenesis. Genes Dev. 21, 3232–3237. Wheater, M.J., Johnson, P.W.M., and Blaydes, J.P. (2010). The role of MNK proteins and eIF4E phosphorylation in breast cancer cell proliferation and survival. Cancer Biol. Ther. 10, 728–735. Wu, M., Liu, Y., Di, X., Kang, H., Zeng, H., Zhao, Y., Cai, K., Pang, T., Wang, S., Yao, Y., and Hu, X. (2013). EIF4E over-expresses and enhances cell proliferation and cell cycle progression in nasopharyngeal carcinoma. Med. Oncol. 30, 400. Xu, J., Chen, A., Joy, J., Xavier, V.J., Ong, E.H., Hill, J., and Chai, C.L. (2013). Rational design of resorcylic acid lactone analogues as covalent MNK1/2 kinase inhibitors by tuning the reactivity of an enamide Michael acceptor. ChemMedChem 8, 1483–1494. Yellen, P., Chatterjee, A., Preda, A., and Foster, D.A. (2013). Inhibition of S6 kinase suppresses the apoptotic effect of eIF4E ablation by inducing TGFb-dependent G1 cell cycle arrest. Cancer Lett. 333, 239–243. Zhang, M., Fu, W., Prabhu, S., Moore, J.C., Ko, J., Kim, J.W., Druker, B.J., Trapp, V., Fruehauf, J., Gram, H., et al. (2008). Inhibition of polysome assembly enhances imatinib activity against chronic myelogenous leukemia and overcomes imatinib resistance. Mol. Cell. Biol. 28, 6496–6509. Zhang, J., Yang, P.L., and Gray, N.S. (2009). Targeting cancer with small molecule kinase inhibitors. Nat. Rev. Cancer 9, 28–39.

Serafimova, I.M., Pufall, M.A., Krishnan, S., Duda, K., Cohen, M.S., Maglathlin, R.L., McFarland, J.M., Miller, R.M., Fro¨din, M., and Taunton, J. (2012). Reversible targeting of noncatalytic cysteines with chemically tuned electrophiles. Nat. Chem. Biol. 8, 471–476.

Zhou, F.F., Yan, M., Guo, G.F., Wang, F., Qiu, H.J., Zheng, F.M., Zhang, Y., Liu, Q., Zhu, X.F., and Xia, L.P. (2011). Knockdown of eIF4E suppresses cell growth and migration, enhances chemosensitivity and correlates with increase in Bax/ Bcl-2 ratio in triple-negative breast cancer cells. Med. Oncol. 28, 1302–1307.

Sherman, W., Day, T., Jacobson, M.P., Friesner, R.A., and Farid, R. (2006). Novel procedure for modeling ligand/receptor induced fit effects. J. Med. Chem. 49, 534–553.

Zuccotto, F., Ardini, E., Casale, E., and Angiolini, M. (2010). Through the ‘‘gatekeeper door’’: exploiting the active kinase conformation. J. Med. Chem. 53, 2681–2694.

12 Chemistry & Biology 21, April 24, 2014 ª2014 Elsevier Ltd All rights reserved

MAP kinase-interacting kinases--emerging targets against cancer.

Mitogen-activated protein kinase (MAPK)-interacting kinases (Mnks) regulate the initiation of translation through phosphorylation of eukaryotic initia...
4MB Sizes 2 Downloads 3 Views